Sunteți pe pagina 1din 185

1521-0081/65/1/315–499$25.00 http://dx.doi.org/10.1124/pr.112.

005660
PHARMACOLOGICAL REVIEWS Pharmacol Rev 65:315–499, January 2013
Copyright © 2013 by The American Society for Pharmacology and Experimental Therapeutics

ASSOCIATE EDITOR: ARTHUR CHRISTOPOULOS

Strategies to Address Low Drug Solubility


in Discovery and Development
Hywel D. Williams, Natalie L. Trevaskis, Susan A. Charman, Ravi M. Shanker, William N. Charman,
Colin W. Pouton, and Christopher J. H. Porter
Drug Delivery, Disposition and Dynamics (H.D.W., N.L.T., W.N.C., C.J.H.P.), Centre for Drug Candidate Optimisation (S.A.C.),
and Drug Discovery Biology (C.W.P.), Monash Institute of Pharmaceutical Sciences, Monash University, Parkville,
Victoria, Australia; and Pfizer Global Research and Development, Groton Laboratories, Groton, Connecticut (R.M.S.)

Downloaded from pharmrev.aspetjournals.org at Queens University Medical Library on October 24, 2013
Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
I. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
A. What Is Low Drug Solubility? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 320
B. Determinants of Aqueous Solubility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323
1. Ideal Versus Nonideal Solubility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323
C. Hydrophobic or Lipophilic Drug Candidates? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 324
D. Solubility of Electrolytes, Weak Electrolytes, and Nonelectrolytes . . . . . . . . . . . . . . . . . . . . . . . 324
E. Solubility and Dissolution Rate. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 325
F. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 326
II. In Vitro Complexities of Working with Poorly Water-Soluble Drugs . . . . . . . . . . . . . . . . . . . . . . . . . 326
A. Drug Precipitation, Adsorption, Binding, and Complexation in In Vitro Assays . . . . . . . . . . 326
B. Changes to Thermodynamic Activity Resulting from Complexation, Binding,
or Solubilization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 327
III. In Vivo Assessment of Poorly Water-Soluble Compounds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 328
A. Parenteral Administration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 329
1. Complexities with Parenteral Administration of Poorly Water-Soluble Drugs . . . . . . . . . 329
2. Parenteral Formulation Approaches for Poorly Water-Soluble Drugs . . . . . . . . . . . . . . . . . 329
B. Oral Administration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331
1. Formulations to Support Drug Discovery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331
2. Use of Enabling Formulations to Promote Oral Absorption . . . . . . . . . . . . . . . . . . . . . . . . . . 333
3. Preclinical Toxicology Formulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333
4. Development of Clinical Formulations for Poorly Water-Soluble Drugs . . . . . . . . . . . . . . . 334
IV. Buffers and Salt Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 334
A. Solution Behavior of Weak Electrolytes and Their Salts. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 335
1. Ionic Equilibria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 335
2. pH Solubility Relationships for Weak Electrolytes and Salts of Weak Electrolytes . . . . 335
3. Factors Affecting Salt Formation at pHmax . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337
4. Determinants of Salt Solubility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 338
a. pHmax . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 338
b. Choice of Counterion to Maximize Salt Solubility. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 338
c. The Effect of Common Ions on Salt Solubility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
d. Effect of Organic Solvents on Salt Solubility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
B. pH adjustment Strategies for Addressing Low Drug Solubility . . . . . . . . . . . . . . . . . . . . . . . . . . 342
1. Buffered Systems Used in Parenteral Formulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
2. Impact of Cosolubilizers and Electrolytes on pH-Mediated Solubilization . . . . . . . . . . . . . 343
a. pH Adjustment and Cosolvents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343
b. pH Adjustment and Strong Electrolytes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 344
c. pH Adjustment and Surfactants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 344

Address correspondence to: Christopher J. H. Porter, Drug Delivery, Disposition and Dynamics, Monash Institute of Pharmaceutical
Sciences, Monash University, 381 Royal Parade, Parkville, VIC, 3052, Australia. E-mail: chris.porter@monash.edu
dx.doi.org/10.1124/pr.111.005660.

315
316 Williams et al.

d. pH Adjustment and Cyclodextrins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 344


3. Effects of Dilution on Drug Solubilization by pH Adjustment . . . . . . . . . . . . . . . . . . . . . . . . 345
a. Methods for Assessing Precipitation Potential for Buffered Parenteral Formulations . 346
4. Buffer Systems in Nonparenteral Formulations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 346
C. The Use of Salts to Address Low Aqueous Solubility in Parenteral Formulations . . . . . . . . 347
D. The Use of Salt Forms to Address Low Aqueous Solubility in Oral Formulations . . . . . . . . 347
1. The Use of Pharmaceutical Salts to Enhance Dissolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . 347
a. Effect of Self-Buffering on Salt Dissolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348
b. Effect of pH Changes on Drug Supersaturation and Precipitation . . . . . . . . . . . . . . . . . 349
c. Potential for Salt Conversion to Un-Ionized Drug/Hydrates/Other Salt Forms In Situ 350
d. Effect of Common Ions on Dissolution in the Gastrointestinal Tract . . . . . . . . . . . . . . 352
2. Physical Properties of Pharmaceutical Salts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 353
3. Potential Toxicity of Pharmaceutical Counterions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 353
E. Feasibility of Salt Formation and Salt Selection Strategies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 354
1. Salt Formation Feasibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 354
2. Salt-Screening Strategies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 355
F. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 356
V. Optimization of Crystal Habit: Polymorphism and Cocrystal Formation . . . . . . . . . . . . . . . . . . . . . 357
A. Polymorphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 358
1. Crystal Packing, Polymorphism, and Phase Transformations . . . . . . . . . . . . . . . . . . . . . . . . 358
2. Effect of Polymorphism on Drug Solubility, Dissolution Rate, and Oral Absorption . . . 359
B. Cocrystals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 360
1. Cocrystals as a Mechanism of Enhanced Drug Solubility and Dissolution . . . . . . . . . . . . 360
2. Solubility Assessment of Cocrystals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 362
3. Solubility Advantages of Cocrystals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 363
4. Design and Preparation of Cocrystals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 365
5. Recent Examples of Pharmaceutical Cocrystals for Improving Solubility
and Bioavailability. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 366
C. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367
VI. Cosolvents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367
A. Rationale for the Use of Cosolvents for the Solubilization of Poorly Water-Soluble Drugs 367
B. Commonly Used Organic Cosolvents in Parenteral Drug Delivery . . . . . . . . . . . . . . . . . . . . . . . 368
C. Solubilization by Cosolvents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 369
1. Cosolvent Effects on Solubility Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 369
2. Predicting Solubility in Cosolvent Systems: The Log-Linear Solubility Model . . . . . . . . . 370
D. Impact of Solubilizers and Electrolytes on Cosolvent-Mediated Drug Solubilization . . . . . . 372
1. Cosolvents and pH Adjustment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 372
2. Cosolvents and Surfactants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 372
3. Cosolvents and Cyclodextrins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 372
4. Cosolvents and Strong Electrolytes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 373

ABBREVIATIONS: ABT-229; 8,9-anhydro-40-deoxy-39-N-desmethyl-39-N-ethylerythromycin B-6,9-hemiaceta; AMG 517, N-(4-[6-(4-trifluoromethyl-


phenyl)-pyrimidin-4-yloxy]-benzothiazol-2-yl)-acetamide; AUC, area under the curve; BCRP, breast cancer–resistant protein; BCS, Biopharma-
ceutical Classification System; CD, cyclodextrin; CMC, critical micelle concentration; CNT, classical nucleation theory; CRA13,
naphthalen-1-yl(4-(pentyloxy)naphthalen-1-yl)methanone; DG, diglyceride; DMA, dimethylacetamide; DMSO, dimethyl sulfoxide; DSC,
differential scanning calorimetry; FA, fatty acid; FABP, fatty acid-binding protein; FaSSGF, fasted-state simulated gastric fluid;
FaSSIF, fasted-state simulated intestinal fluid; FATP, fatty acid transport protein; FeSSIF, fed-state simulated intestinal fluid; FTIR,
Fourier transform infrared spectroscopy; GI, gastrointestinal; HDL, high-density lipoprotein; HPC, hydroxypropyl cellulose; HLB,
hydrophilic-lipophilic balance; HPMC, hydroxypropyl methylcellulose; HPH, high-pressure homogenization; HPMCAS, hydroxypropyl
methylcellulose acetate succinate; K-832, 2-benzyl-5-(4-chlorophenyl)-6-[4-(methylthio)phenyl]-2H-pyridazin-3-one; L-883555, N-cyclopropyl-
1-{3-[6-(1-hydroxy-1-methylethyl)-1-oxidopyridin-3-yl]phenyl}-1,4-dihydro-[1,8]naphthyridin-4-one 3-carboxamide; LBF, lipid-based for-
mulations; LC, long chain; LCQ789, 5-(4-chlorophenyl)-1-phenyl-6-(4-(pyrazin-2-yl)phenyl)-1H-pyrazolo[3,4-d]pyrimidin-4(5H)-one;
LDL, low-density lipoprotein; LFCS, Lipid Formulation Classification System; LPC, lysophosphatidylcholine; MC, medium chain; mdr
or MDR, multi-drug resistant; MG, monoglyceride; MRP, multiresistance protein; NMR, nuclear magnetic resonance; NSC-639829, N-
[4-(5-bromo-2-pyrimidyloxy)-3-methylphenyl]-(dimemethylamino)-benzoylphenylurea; OZ209, cis-adamantane-2-spiro-39-89-(aminomethyl)-
19,29,49-trioxaspiro[4.5]decane mesylate; PEG, polyethylene glycol; PG-300995, 2-(2-thiophenyl)-4-azabenzoimidazole; P-gp, P-glycoprotein; PL,
phospholipid; PPI, polymer precipitation inhibitor; PVP, polyvinylpyrrolidone; RPR200765, {t-2-[4-(4-fluoro-phenyl)-5-pyridin-4-yl-1H-imidazol-
2-yl]-5-methyl-[1,3]dioxan-r-5-yl}-morpholin-4-yl-methanone;; SCF, super critical fluids; SEDDS, self-emulsifying drug-delivery systems; SD,
solid dispersion; SGF, simulated gastric fluid; SLN, solid lipid nanoparticle; TG, triglyceride; TPGS, D-a-tocopheryl polyethylene glycol succinate;
TRL, triglyceride-rich lipoprotein; UWL, unstirred water layer; XRPD, X-ray powder diffraction.
Strategies for Low Drug Solubility 317

E. Effects of Dilution on Drug Solubilization by Cosolvents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 374


F. Potential Pharmacological Properties of Cosolvents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 374
1. In Vitro Assessment of Cosolvent Toxicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 374
2. Lytic Effects of Commonly Used Cosolvents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 375
G. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 375
VII. Surfactants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 375
A. Rationale for the Use of Surfactants to Enhance the Solubilization of Poorly
Water-Soluble Drugs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 375
B. Commonly Used Nonionic Surfactants in Drug Delivery. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 377
C. Solubilization by Surfactants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 378
1. Micelle Formation and Drug Solubilization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 378
2. Quantification of Micellar Solubilization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 379
3. Effect of Surfactant Type and Structure on Drug Solubilization . . . . . . . . . . . . . . . . . . . . . . 380
4. Effect of Temperature on Drug Solubilization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 380
D. Impact of Cosolubilizers and Electrolytes on Surfactant-Mediated Drug Solubilization . . . 381
1. Surfactants and Cosolvents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 381
2. Surfactants and pH Adjustment. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 381
3. Surfactant Mixtures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 382
4. Surfactants and Cyclodextrins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 383
5. Surfactants and Strong and Weak Electrolytes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 383
E. Effects of Dilution on Drug Solubilization by Surfactants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 383
F. Potential Pharmacological Properties of Nonionic Surfactants . . . . . . . . . . . . . . . . . . . . . . . . . . . 384
1. Toxicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 384
2. Pharmacokinetic Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 386
G. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 387
VIII. Cyclodextrins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 387
A. Rationale for the Use of Cyclodextrins in the Solubilization of Poorly Water-Soluble Drugs. . 387
B. Drug-Cyclodextrin Inclusion Complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 389
1. Binding Equilibria between a Drug and Cyclodextrin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 389
2. Measurements of Drug-Cyclodextrin Binding Constants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 390
3. Secondary Equilibria: Micelles and Cyclodextrin Aggregate Formation . . . . . . . . . . . . . . . 392
C. Drug Factors Affecting Complexation by Cyclodextrins. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393
1. Size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393
2. Polarity and Charge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393
3. Presence of Counterions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 395
D. Impact of Cosolubilizers and Excipients on Cyclodextrin-Mediated Drug Solubilization . . 395
1. Cyclodextrins and Water-Soluble Polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 395
2. Cyclodextrins and Cosolvents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 396
3. Cyclodextrins and Surfactants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 396
E. In Vivo Utility of Drug-Cyclodextrin Complexes in Parenteral Formulations . . . . . . . . . . . . . 396
1. Effect of Cyclodextrins on Drug Pharmacokinetics after Intravenous Administration . 396
2. Effects of Dilution on Drug Solubilization by Cyclodextrins . . . . . . . . . . . . . . . . . . . . . . . . . . 397
3. Potential Toxicological Properties of Parenterally Administered Cyclodextrins. . . . . . . . 397
F. In Vivo Utility of Drug-Cyclodextrin Complexes in Oral (Nonparenteral) Formulations . . 398
1. Effect of Cyclodextrins on Oral Drug Bioavailability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 398
2. Oral Administration of Physical Mixtures Versus Isolated
Drug-Cyclodextrin Complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 399
3. Using Cyclodextrins to Stabilize Amorphous Drug . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 399
4. Cyclodextrin Complexation as a Limitation to Oral Bioavailability . . . . . . . . . . . . . . . . . . . 400
5. Potential Toxicological Properties of Orally Administered Cyclodextrins. . . . . . . . . . . . . . 401
6. Effect of Cyclodextrins on Membrane Transport and Drug Metabolism . . . . . . . . . . . . . . . 402
G. Manufacture of Cyclodextrin Formulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 402
1. Parenteral Formulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 402
2. Solid Drug-Cyclodextrin Complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 402
H. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 402
IX. Particle Size Reduction Strategies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 403
A. Particle Size Effects on Dissolution, Solubility, and In Vivo Performance . . . . . . . . . . . . . . . . 404
318 Williams et al.

1. Effect of Particle Size on Dissolution Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 404


2. Effect of Particle Size on Saturated Solubility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 405
3. Effect of Particle Size and Shape on the Diffusional Layer Thickness . . . . . . . . . . . . . . . . 405
B. Common Methods to Reduce Particle Size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 406
1. “Top-Down” Particle Size Reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 406
a. Pearl Milling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 406
b. High-Pressure Homogenization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 407
2. “Bottom-up” Nanoparticle Assembly . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 408
a. Controlled Crystallization Via Solvent Shift . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 409
b. Precipitation after Solvent Evaporation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 410
C. Oral and Parental Delivery of Formulations Containing Nanosized Drug Particles. . . . . . . 411
1. Oral Delivery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 411
2. Nanosuspensions for Parenteral Delivery of Poorly Water-Soluble Drugs . . . . . . . . . . . . . 411
D. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 415
X. Solid Dispersions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 416
A. Classification of Solid Dispersions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 417
1. Nontraditional Solid Dispersions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 419
B. Mechanisms by which Solid Dispersions Enhance Dissolution Rate and
Oral Bioavailability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 419
1. Reduced Particle Size and Enhanced Drug Wetting and Solubilization . . . . . . . . . . . . . . . 419
2. Administration of Drug in the Amorphous Form. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 421
3. Maintenance of Supersaturation after Drug Dissolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 421
a. The Impact of Dissolution Rate and Supersaturation Ratio on
Precipitation Inhibition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 423
b. Mechanisms by which Polymers Maintain Supersaturation in Solution . . . . . . . . . . . 423
c. Screening for Potential Polymer Precipitation Inhibitors. . . . . . . . . . . . . . . . . . . . . . . . . . 425
C. Recent Examples of Bioavailability Enhancement Using Solid Dispersions . . . . . . . . . . . . . . 426
D. Role of the Carrier and the Drug-Carrier Ratio in Dictating Drug Release Kinetics
from Solid Dispersions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 427
E. Common Methods to Manufacture Solid Dispersions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 429
1. Melting/Fusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 429
2. Solvent Evaporation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 429
3. Solvent Evaporation Using Supercritical Fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 430
4. Electrospinning and Microwave Irradiation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 431
F. Physical Stability of Amorphous Solid Dispersions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 431
1. Structural Changes at the Glass Transition Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . 431
2. Glass-Forming Potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 432
3. Polymer Effects on the Glass Transition Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 433
4. Moisture Effects on Glass Transition Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 434
G. Factors Affecting Drug Crystallization from Solid Dispersions. . . . . . . . . . . . . . . . . . . . . . . . . . . 434
1. Drug-Polymer Intermolecular Association . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 435
2. Inhibition of Crystal Nucleation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 435
3. Molecular Mobility, Strength, and Fragility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 436
4. Limitations to the Use of Thermal Analysis to Estimate Molecular Mobility. . . . . . . . . . 439
5. Drug-Polymer Miscibility and Relationship to Drug Loading . . . . . . . . . . . . . . . . . . . . . . . . . 439
H. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 444
XI. Lipid-Based Formulations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 445
A. Mechanisms of Bioavailability Enhancement by Lipid-Based Formulations . . . . . . . . . . . . . . 446
1. Stimulation of Intestinal Lipid Absorption and Transport Pathways . . . . . . . . . . . . . . . . . 446
2. Enhanced Drug Dissolution and Solubilization in the Intestinal Lumen . . . . . . . . . . . . . . 450
3. Enhanced Intestinal Permeability and Inhibition of Intestinal Efflux and
First-Pass Metabolism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 452
a. Correlation of In Vitro Effects of Lipid-Based Formulation Excipients on
Permeability with Changes to In Vivo Exposure?. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 457
4. Promotion of Lymphatic Drug Transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 459
B. Design and Formulation of Lipid-Based Formulations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 461
1. Lipid-Based Fomulations for Parenteral Administration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 461
Strategies for Low Drug Solubility 319

2. Lipid-Based Formulations for Oral Administration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 463


3. The Lipid Formulation Classification System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 464
4. Selection of Lipid-Based Formulation Excipients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 465
a. Solvent Capacity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 466
b. Mutual Miscibility. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 466
c. Toxicity/Irritancy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 466
d. Purity and Chemical Complexity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 467
e. Capsule Compatibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 467
f. Polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 468
C. Assessment of Lipid-Based Formulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 468
1. In Vitro Dispersion Testing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 468
2. In Vitro Digestion Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 468
3. In Vivo Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 470
D. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 471
XII. Emerging Strategies for Improving the Aqueous Solubility of Poorly Water-Soluble Drugs . . . 472
A. Drug Adsorption to Microporous Adsorbents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 472
B. Solid Lipid Nanoparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 473
XIII. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 474
Acknowledgments. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 475
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 476

Abstract——Drugs with low water solubility are pre- and review the approaches and strategies that can be
disposed to low and variable oral bioavailability and, taken to facilitate compound progression. In particular,
therefore, to variability in clinical response. Despite we address the fundamental principles that underpin the
significant efforts to “design in” acceptable developabil- use of strategies, including pH adjustment and salt-form
ity properties (including aqueous solubility) during lead selection, polymorphs, cocrystals, cosolvents, surfac-
optimization, approximately 40% of currently marketed tants, cyclodextrins, particle size reduction, amorphous
compounds and most current drug development candi- solid dispersions, and lipid-based formulations. In each
dates remain poorly water-soluble. The fact that so many case, the theoretical basis for utility is described along
drug candidates of this type are advanced into develop- with a detailed review of recent advances in the field.
ment and clinical assessment is testament to an increas- The article provides an integrated and contemporary
ingly sophisticated understanding of the approaches discussion of current approaches to solubility and
that can be taken to promote apparent solubility in the dissolution enhancement but has been deliberately
gastrointestinal tract and to support drug exposure after structured as a series of stand-alone sections to allow
oral administration. Here we provide a detailed com- also directed access to a specific technology (e.g., solid
mentary on the major challenges to the progression of dispersions, lipid-based formulations, or salt forms)
a poorly water-soluble lead or development candidate where required.

I. Introduction for highly lipophilic, poorly water-soluble drug candidates


to access and interact with the target.
High hydrophobicity and intrinsically low water solu-
These drivers ultimately bias the identification of
bility are increasingly common characteristics of hits,
poorly water-soluble “hits” during early drug screens.
leads, development candidates, and ultimately marketed
drugs. Many hypotheses have been put forward as to Poor water solubility is a significant risk factor in low
why these trends have emerged, and the true explana- oral absorption because drug molecules must, in most
tion is clearly multifaceted. The application of combi- cases, be in solution to be absorbed, and oral bioavail-
natorial chemistries to generate large chemical libraries ability is usually a required characteristic in a target
and the common application of high-throughput screen- product profile of an orally administered medicine. As
ing modalities, often in nonaqueous media (or mixed- such, medicinal chemistry strategies during lead opti-
solvent media), have probably played a role. The desire mization typically seek to modify physicochemical pro-
for increased potency, coupled with the realization that perties (including solubility) such that drug leads have
receptor binding is mediated, at least in part, by more developable characteristics. Many decision gates,
hydrophobic interactions, further magnifies the likelihood or idealized character panels, are used to identify and
that drug candidates will have limited aqueous solubility. reject drug candidates with inappropriate developability
Finally, the quest to explore unprecedented drug targets, properties and, subsequently, to synthetically modify
some of which are associated with intracellular signaling structures to improve physicochemical characteristics.
pathways, lipid processing architecture, or highly lipo- Perhaps the best known of these is Chris Lipinski’s “rule
philic endogenous ligands, only amplifies the requirement of 5” (Lipinski et al., 1997), but there are many others.
320 Williams et al.

In all cases, however, at least moderate water solubility This review is structured to provide an initial and
is usually a focus. relatively brief introduction to the determinants of low
Nonetheless, even with contemporary medicinal water solubility to provide the theoretical basis for
chemistry programs and increasingly sophisticated the approaches that might be taken to address solu-
lead optimization strategies, it is apparent that for bility challenges. We subsequently present summary
some targets, reducing lipophilicity and increasing sections that outline potential strategies for addressing
water solubility will result in an unacceptable reduction solubility issues, first in vitro and second in vivo, with
in potency. In spite of attempts to circumvent solubility the latter section addressing both parenteral and oral
problems, approximately 40% of currently marketed administration. Subsequently, we provide a compre-
drugs (Fig. 1) and up to 75% of compounds currently hensive review of the technologies that can be used to
under development have been suggested to be poorly promote solubility or dissolution. These latter sections
water-soluble (Di et al., 2009, 2011). Furthermore, the are necessarily dense and are intentionally separated
problems of low water solubility do not seem to be from the higher-level strategy summaries provided in
diminishing and may well be increasing (Takagi et al., the introduction. The technology overviews provide
2006). Low water solubility therefore continues to be a reference source for the summaries that precede them.
a challenge to successful drug development. A schematic representation of each of these formulation
This review seeks to provide an overview of the approaches is provided in Fig. 2 and is intended to
strategies that may be taken to address the problems highlight the variety of proven strategies available to
of low solubility during drug discovery and develop- those working with poorly water-soluble drugs.
ment and includes a comprehensive review of formu-
lation approaches to support the clinical development A. What Is Low Drug Solubility?
of oral and parenteral drug products for poorly water- Although low water solubility of drug candidates
soluble drugs. In addition, even when lead optimiza- presents varied and significant challenges throughout
tion is successful in increasing aqueous solubility, early drug discovery and development, the greatest concern
preclinical studies are still required with less than is generally the risk of reduced and variable absorp-
optimal leads to provide the data sets to allow informed tion after oral administration. The value at which
progression or rejection; therefore, we also address the limited solubility begins to impact absorption is
strategies that might be used when dealing with the difficult to state definitively since it is dependent on
complexities of low solubility during the discovery a number of other system variables, including drug
phase. In the former case, at least for oral drug products, permeation, dose, and the environment present
the market typically dictates the need for traditional within the gastrointestinal (GI) tract. To understand
solid dosage forms (e.g., capsules, tablets). In the latter, these variables, and therefore, the factors that impact
where studies are usually conducted in small animals the required solubility for a drug candidate, it is
(rodents), liquid formulations are often used to allow instructive first to appreciate the drivers of flux across an
oral dosing, and therefore a slightly different ap- absorptive membrane since these in turn will determine
proach must be taken. whether the drug dose can be absorbed over the timescale
available.
Assuming appropriate chemical and metabolic sta-
bility and an absence of transporter or carrier-mediated
processes, flux (F) across an absorptive membrane is the
product of the concentration gradient and the perme-
ability across the membrane and is described as follows:
D×K×A
F5 ðCm 2 C0 Þ ð1Þ
h
where Cm is the drug concentration immediately
adjacent to the membrane, C0 is the drug concentration
on the abluminal side of the membrane, K is the par-
tition coefficient between the aqueous solution overlay-
ing the membrane and the membrane itself, h is the
width of the diffusion layer, D is the diffusion coefficient
Fig. 1. A comparison of the distribution of solubilities for the top 200 oral of the drug in the membrane, and A is the surface area.
drug products in the United States (US), Great Britain (GB), Spain, and
Japan and from the World Health Organization (WHO) Essential Drug Assuming that the drug concentration on the abluminal
List. Very soluble drugs: over 1000 mg/ml; freely soluble drugs: 100–1000 side of the absorptive membrane is low relative to the
mg/ml; soluble drugs: 33–100 mg/ml; sparingly soluble drugs: 10–33 mg/ml;
slightly soluble drugs: 1–10 mg/ml; very slightly soluble drugs: 0.1–1 mg/ml; concentration at the absorptive site (i.e., Cm . . C0), eq.
practically insoluble drugs: ,0.1 mg/ml. Adapted from Takagi et al. (2006). 1 can be simplified to the following:
Strategies for Low Drug Solubility 321

Fig. 2. Schematic diagram illustrating the common strategies currently used (and discussed in this review) to address low drug solubility in drug
discovery and development.

D×K×A originally derived by Johnson and Swindell (1996) and


F5 Cm ð2Þ
h further applied by Curatolo (1998) and Lipinski (2000):
The values for D, K, and h are typically fixed for MAD 5 Cs × kabs × SIWV × SITT ð5Þ
a particular system and are used to define the
permeability coefficient (P) where where Cs is the solubility (mg/ml) at pH 6.5 (represent-
ing the pH of the small intestine); kabs is the rate
D×K constant (h21) for intestinal absorption (which is
P5 ð3Þ
h related to the permeability); SIWV is the small
intestinal water volume (in milliliters), which is
In the absence of supersaturation, the maximum
typically assumed to be ;250 ml (the volume of fluid
concentration that can be attained at the surface of an
assumed to be present in the fasted GI tract after
absorptive membrane is equivalent to the equilibrium
a glass of water has been drunk when taking
solubility (Cs) of the drug, and therefore the maximum
medication orally); and SITT is the small intestinal
flux (per unit area) (F’) is the product of solubility and
transit time (min) of ;270 min (4.5 h). Rearranging
permeability:
this relationship provides an expression for the
F9 5P × Cs ð4Þ necessary or target solubility for a given dose and kabs
(or permeability) and provides an initial indication as
Appreciation of this relationship illustrates that to whether solubility is likely to limit oral absorption.
knowledge of the solubility alone is insufficient to This concept is shown graphically in Fig. 3, the data
anticipate whether solubility will limit flux (or absorp- from which are taken from a seminal review that
tion) since flux is also a function of permeability. To shows the theoretical required solubility to provide
some extent, therefore, low solubility can be offset by good oral absorption for drugs with projected doses
high permeability; similarly, if permeability is low, the ranging from 0.1 to 10 mg/kg and permeabilities
requirements for solubility to generate appropriate ranging from low to high. At one end of the spectrum,
flux increase. highly potent drugs for which the dose is low and the
A well recognized approach applied in early drug membrane permeability is high have relatively low
discovery for estimating the required solubility and solubility requirements to achieve good oral absorp-
permeability needed to achieve good oral absorption is tion. At the other extreme, low-potency drugs for which
the concept of a maximum absorbable dose (MAD), the dose is high and the permeability is low need
322 Williams et al.

permeability properties. Originally proposed to provide


a scientific basis for biowaivers based on a correlation of
in vitro drug dissolution and in vivo drug absorption,
this classification system has found much broader
applicability across many areas of drug discovery and
development. According to the BCS, class I molecules
are those having both high solubility and high perme-
ability (and therefore likely few problems with oral
absorption); class II compounds are those that have low
solubility and high permeability (where solubility is the
primary limitation to absorption); class III compounds
have high solubility but low permeability (where
absorption is limited by membrane permeation and
not solubility); and class IV compounds are those in
which both poor solubility and poor permeability limit
Fig. 3. The relationship between the projected dose and the required drug absorption. The focus of the current review is
aqueous solubility for low, medium, and high permeability compounds. therefore BCS class II compounds, which often exhibit
Permeability is indicated by the magnitude of the absorption rate
constant (kabs), where low permeability, kabs 5 0.1; medium 5 1.0, high 5
solubility-limited absorption. Class IV compounds are
10 h21. From this analysis, moderately permeable compounds (kabs 5 also relevant, although they have additional prob-
1.0 h21), with projected potencies of 1.0 mg/kg, require aqueous lems associated with low permeability.
solubilities of ;52 mg/ml. Adapted from Lipinski (2000).
Drug dose is also an important factor in the BCS
because highly soluble drugs are defined as those in
considerably higher solubility for good oral absorption which the highest dose will dissolve in 250 ml over
(by several orders of magnitude in this example). As the pH range of the GI tract (i.e., pH 1–6.8). Several
a broad initial estimation, and assuming moderate examples of dose, solubility, and volume requirements
potency (1.0 mg/kg) and moderate permeability (kabs 5 (taken from Amidon et al., 1995) are shown in Table 1.
1.0 h21), this approach indicates that where aqueous As with the MAD, these calculations are not designed to
solubilities are ,50 mg/ml, problems associated with be definitive but rather illustrate that low-dose com-
low water solubility might be anticipated. It is clear, pounds, such as digoxin, can have good absorption and
however, that the required solubility to support drug bioavailability, even when GI solubility is low, whereas
absorption must be evaluated in light of both the the absorption of high-dose compounds, such as griseo-
potency (or dose) and the permeability characteristics. fulvin, is more often low, variable, and highly formula-
It is also clear that at stages during the development tion dependent as a result of their solubility limitations.
pathway, in particular during preclinical toxicity A complication of the BCS definition of high
testing, exposure at doses considerably in excess of solubility is that the highest strength dose must be
the predicted clinical dose will be required, magnifying soluble in 250 ml of water at all pH values that might
the need for solubility support. be encountered in the GI tract. Therefore, drugs may
A further application of the solubility-permeability be classified as class II even though they have good
relationship to oral drug absorption is the Biophar- solubility at one end of this pH range. For example,
maceutics Classification System (BCS) (Fig. 4), many weak acids have low solubility at pH 1 and are
originally developed by Amidon et al. (1995), with strictly classified as BCS class II compounds but are
subsequent variations by others (Wu and Benet, 2005; quite soluble at intestinal pH (pH 6–7) and in many
Butler and Dressman, 2010; Chen et al., 2011). The cases do not exhibit solubility-limited absorption. As
principles of the BCS are well described elsewhere such, a BCS class II designation does not always
(Amidon et al., 1995; Yu et al., 2002; Dahan et al., 2009), dictate that solubility will be a limitation to absorption;
but in brief, the BCS allows classification of drug rather, compounds in this BCS class are more likely to
molecules as a function of their solubility and be solubility limited than are those in class I. Indeed,

TABLE 1
Examples of dose, solubility, and volume requirements for a range of poorly water-soluble drugs
Volume Required for
Drug Dose Solubility Complete Solubilization Absorption

mg mg/ml ml
Piroxicam 20 0.007 2857 Low, variable
Digoxin 0.5 0.024 21 Good
Griseofulvin 500 0.015 33,333 Low, variable
Chlorthiazide 500 0.78 636 Reasonable
From Amidon et al. (1995).
Strategies for Low Drug Solubility 323

Fig. 4. Diagrammatic representation of the BCS, which classifies drugs


according to their permeability and solubility properties. Compounds are Fig. 5. Three essential steps are required for a solute drug molecule to be
defined as high solubility when the quantity of drug that is present in displaced from a solid particle and to enter solution. Step 1: A single
the highest strength immediate release dose form is soluble in 250 ml of solute molecule is removed from the crystal lattice; energy is required in
water across the likely range of gastrointestinal pH (1.2–7.5). Highly this step to overcome solute-solute interactions in the solid state. Step 2:
permeable compounds are defined as those that are .90% absorbed A void is created within the solvent to accommodate the solute molecule.
or where permeability as assessed by in vitro or in vivo methods is Although this step also requires energy, it is likely to be considerably
equivalent to or higher than that of a reference compound that is 90% lower than the energy required in step 1. Step 3: The solute molecule
absorbed. Adapted from Amidon et al. (1995). inserts into the solvent, forming solute-solvent interactions. Simplisti-
cally, if the energy released from the solute-solvent interactions (i.e., step
3) is greater than the energy required for steps 1 and 2, solubility is
favored.
a more practical description of BCS class II may be
drugs that do not have high solubility rather than
those that have low solubility since the classification compared with nonplanar molecules. The melting point
system specifically identifies compounds that fall into provides a reasonable indication of the strength of in-
class I and all those with solubilities below this fall termolecular solute-solute interactions in the solid
inherently into class II (or class IV if permeability is material. Second, a void must be created in the solvent
also low) (Fig. 4). that is sufficient to accommodate the abstracted solute
The preceding discussion serves to illustrate further molecule. Since intermolecular forces in the liquid
that low solubility is a somewhat arbitrary concept state are much lower than those in the solid state, the
when assessing the likelihood that solubility will limit energy required to create a void in the solvent is low
drug absorption and that additional knowledge of the and is usually ignored when assessing energy changes
likely dose and membrane permeability is inevitably during dissolution. Finally, the solute molecule is
required to put a solubility value into an appropriate inserted into the solvent void. For molecules with some
context. affinity for a polar solvent such as water, this process is
energetically favorable and therefore drives drug solu-
B. Determinants of Aqueous Solubility bility. This last concept is the rationale behind the often
A detailed description of the thermodynamic deter- quoted maxim “like dissolves like.” This is largely true,
minants of drug solubility in aqueous media is beyond but it captures only half the story since it ignores the
the scope of the current discussion; for more informa- impact of changes to solid-state properties on solubility.
tion, the interested reader is directed to the following Since the energy transitions associated with changes
references: Grant and Higuchi (1990), Yalkowsky (1999), to solvent structure are small, it is apparent that there
and Murdande et al. (2010a). An overview of the broad are two primary determinants of drug solubility: 1) the
physicochemical drivers of drug solubility is war- energy required to overcome the strength of intermo-
ranted, however, since these drivers underpin the lecular forces in the solute solid state and 2) the energy
approaches that can be taken to enhance solubility. generated on the interaction of solute and solvent
Simplistically, the potential for a drug (solute) to pass molecules in solution (solvation)
into solution in an aqueous fluid (solvent) is dictated by For an analogous structural series, solubility is
three separate events, as shown in Fig. 5. First, solute therefore generally lower for molecules that have
molecules must be abstracted from the solid state, higher melting points (stronger attractive forces) and
a process that involves breaking solute-solute bonds. lower affinity for water (poor solvation).
The strength of solute-solute attractive forces in dif- 1. Ideal Versus Nonideal Solubility. Solubility
ferent solids varies significantly and is typically higher theory defines an idealized condition or an ideal solu-
for electrolytes than for nonelectrolytes (since ionic tion, where the intermolecular forces between solute
attractive forces in the solid state are stronger than and solvent are equivalent to those between solute and
nonionic forces), for crystalline solids compared with solute and those between solvent and solvent. Under
amorphous materials, and for planar nonelectrolytes these circumstances, the mixing of solute and solvent
(which pack more effectively in the solid state) molecules in the liquid state results in no net energy
324 Williams et al.

change. Where this is the case, solubility is dependent between poorly water-soluble drugs that are limited by
on the strength of the solute crystal lattice and may be solid-state properties (e.g., the strength of the crystal
defined by lattice), those that are limited by solvation (i.e., solute-
  solvent interactions in solution), and those that are
DHf Tm 2 T
logX5 2 ð6Þ limited by both. In practice, most compounds fall into
2:303R Tm T two categories since almost all poorly water-soluble
compounds have limited affinity for water (i.e., they are
where X is the ideal mole fraction solubility of solute,
hydrophobic) and are therefore solvation-hydration
DHf is the enthalpy of fusion of solute, Tm is solute
limited. Where compounds are hydrophobic and also
melting point, T is the absolute temperature, and R is
show strong intermolecular forces in the solid state,
the gas constant.
they are typically poorly soluble in both aqueous and
In reality, ideal solutions are highly unusual, and
nonaqueous solvents. In contrast, hydrophobic mole-
solution conditions close to ideality exist pharmaceu-
cules, in which solubility is not limited by solid-state
tically only in solutions of highly lipophilic drugs in
properties, often show varying degrees of solubility in
nonaqueous (lipidic) solutions. In contrast, in aqueous
nonaqueous solvents such as lipids (since the molecular
solution, the differences between solute-solute inter-
properties that reduce hydration in aqueous media often
actions and solute-solvent interactions are highly sig-
promote solvation in nonaqueous media). The latter
nificant, leading to nonideal solution behavior and
compounds are therefore both hydrophobic and lipo-
much lower solubility than would otherwise be pre-
philic, with the former simply being hydrophobic. The
dicted. In this case, the difference between ideal and
difference between these two types of molecules can be
nonideal behavior is captured by a correction factor
illustrated using the analogies of “brick dust” for hydro-
termed the activity coefficient (g), where
phobic molecules with poor solubility in all solvents and
  “grease balls” for compounds that are hydrophobic and
DHf Tm 2 T
logX5 2 2 log g ð7Þ lipophilic and show reasonable solubility in lipids (Stella
2:303R Tm T
and Nti-Addae, 2007). This distinction is important
In turn, the activity coefficient is a function of the since the formulation options available for either
molar volume of the solvent (Vs), the volume fraction of hydrophobic or lipophilic compounds differ considerably.
the solute (w1), and the difference in the solubility Simplistically, the increased lipid solubility of lipophilic
parameters for the solvent (d1) and solute (d2): drug candidates allows access to liquid surfactants and
lipid-based delivery technologies that can be filled into
Vs w21 soft gelatin capsules (or sealed hard gelatin capsules),
log g 5 ðd1 2 d2 Þ2 ð8Þ
2:303RT whereas the lack of solubility for hydrophobic molecules
The solubility parameters provide a measure of the in almost all vehicles precludes formulation in anything
cohesive forces in either solute or solvent, and from eq. 8, other than modified solid dosage forms.
it is evident that smaller differences between the D. Solubility of Electrolytes, Weak Electrolytes,
solubility parameters of the solvent and solute give rise and Nonelectrolytes
to lower activity coefficients and, therefore, an increase
in solubility toward ideality (i.e., d1 5 d2) (Rubino, 2002). The solvation properties of drug molecules, and
The solubility equation (eq. 7) captures the determi- therefore a significant part of the driving force for drug
nants of solubility as dictated by changes to the solid- solubility in aqueous media, are highly dependent on the
state properties of a drug (and usually manifest by extent of ionization. The presence of a charged functional
changes to melting point) and the activity coefficient group provides an opportunity for favorable ion-dipole
(reflecting differences between the solubility parame- interactions with polar solvents such as water, which
ters), and it reiterates the importance of solute-solute directly enhance hydration and water solubility. Strong
interactions in the solid state and the need to minimize electrolytes (such as NaCl) that completely dissociate in
differences between solute and solvent properties to water are generally highly water soluble. This is not
maximize solvation (i.e., like dissolves like). These always the case, however, as solubility remains a com-
principles underpin all the approaches to solubility posite function of the energy required to break the
enhancement that are subsequently described in this crystal lattice versus the energy liberated on hydration
review, as each of these approaches either change solid- of the ions formed. In some extreme cases, for example,
state properties or change the nature of the interaction an inorganic salt such as AgCl, the crystal lattice energy
between drug (solute) and solvent molecules. is sufficiently high that aqueous solubility remains low.
As a strong electrolyte, therefore, dissociation of AgCl
molecules in solution is complete, but as a poorly water-
C. Hydrophobic or Lipophilic Drug Candidates? soluble strong electrolyte, solubility is limited.
An understanding of the primary drivers of drug In reality, most drugs are organic materials that are
solubility allows an important distinction to be made either nonelectrolytes (which do not dissociate to form
Strategies for Low Drug Solubility 325

ionic species in water) or weak electrolytes that


dissociate only partially in water such that both un-
ionized solute and the dissociated ions are present in
solution.
The solubility of weak electrolytes is highly de-
pendent on the degree of ionization (dissociation) since
the affinity of the ionized species for water is markedly
higher than that of the un-ionized species. The degree
of dissociation is in turn dependent on the pKa of the
weak electrolyte and the pH of the solution into which
it is dissolving. Simplistically, at pH values above the
pKa of a weak acid and below the pKa of a weak base,
solubility increases significantly as a result of ioniza-
tion. For weak acids and bases with a single ionizable
group, solubility increases by a factor of 10 for every pH
unit away from the pKa (although this trend does not
continue indefinitely and solubility is ultimately Fig. 6. Graphic depicting the process of drug dissolution from a solid
limited by the solubility product of the salts that are drug particle. An unstirred water layer of width (h) is present on the
formed in situ with available counterion; see Section surface of the dissolving solid. A concentration gradient is established
across the unstirred water layer that drives dissolution and results from
IV). Nonetheless, optimization of solution pH is an the difference in drug concentration between that on the surface of the
effective means by which solubility can be enhanced dissolving solid (usually assumed to be the equilibrium solubility of the
drug, Cs) and the concentration in the well stirred bulk (C).
and is commonly used to enhance the solubility
properties of solution formulations for weak electro-
lytes. For solid dosage forms, the principles of pH-
available. For different drugs, and under different
dependent solubility may be manipulated by the
circumstances, either solubility or dissolution rate (or
isolation of a drug (or drug candidate) as an appropri-
both) may be the limiting feature.
ate salt form. The use of pH and salt selection to
The process of drug dissolution from the solid state is
enhance solubility and the dissolution rate is described
summarized in Fig. 6. An unstirred water layer is
in more detail in Section IV. For nonelectrolytes,
present on the surface of every dissolving solid and
solubility behavior is not complicated by the effects of
provides the barrier to drug equilibration with the well
ionization and remains a function of hydration and the
stirred bulk solution. The dissolution rate is therefore
strength of the crystal lattice.
defined by the rate at which drug diffuses across the
E. Solubility and Dissolution Rate unstirred water layer, and the equations that describe
drug dissolution (i.e., the Noyes Whitney Equation, eq.
The solubility of a drug in aqueous solution is
9) are analogous to simple diffusion equations. The rate
a fundamentally important property that affects not
of transfer across the unstirred layer is a function of the
only the potential for drug absorption after oral
concentration gradient across the unstirred layer, the
administration and the ability to administer the drug
width of the diffusion layer (h), the surface area of
parenterally but also the ease of manipulation and
contact of the solid with the dissolution fluid (A), and
testing in the laboratory and during manufacture.
the diffusion rate of the drug in water (D). The
However, drug solubility is an equilibrium measure
concentration gradient in turn is a function of the
and the rate at which solid drug or drug in a formula-
maximum drug concentration at the surface of the
tion passes into solution (i.e., the dissolution rate) is
dissolving solid (drug solubility Cs) and the concentra-
also critically important when the time available for
tion in the well-stirred bulk (C) (Noyes and Whitney,
dissolution is limited. This rate is particularly relevant
1897):
after oral administration since intestinal transit time
is finite and the rate of drug dissolution must sig- dcx D × A
nificantly exceed the rate of transit for absorption to be 5 ðCs 2 CÞ ð9Þ
dt h
maximized. For example, the absorption of a drug with
reasonable solubility may still be poor if the rate of
dissolution is low since the solubility limit may never If we assume sink conditions where the concentra-
be reached during the intestinal transit time. Simi- tion of drug in bulk solution (C) is low relative to the
larly, even where the rate of dissolution is relatively concentration on the surface of the dissolving solid (Cs),
rapid, if the equilibrium solubility is low, the quantity then this relationship collapses to
of drug available in solution for absorption is unlikely
to support rates of flux across the intestine that are dcx D × A
5 Cs ð10Þ
sufficient to absorb the entire drug dose in the time dt h
326 Williams et al.

For most small molecules, the diffusion rate constant Nti-Addae (2007), Di et al. (2009), Keseru and Makara
in water (D) is relatively high and manipulation of (2009), and Ishikawa and Hashimoto (2011).
drug structure does not typically affect D to the extent
that it has a significant impact on dissolution rate.
II. In Vitro Complexities of Working with Poorly
Similarly, whereas the width of the diffusion layer can
Water-Soluble Drugs
be altered by agitation or stirring in vitro, this aspect
cannot be easily manipulated in vivo. Despite the drive to identify “drug-like” leads with
The major determinants of in vivo drug dissolution acceptable physicochemical properties, lead series are
rate are therefore solubility and surface area. Since often plagued by poor aqueous solubility. The basis for
solubility is a function of the strength of the crystal this trend was discussed earlier in this introduction,
lattice and the affinity of the solute (drug) for the but regardless of the source, working with inherently
aqueous environment, three major strategies can be poorly water-soluble compounds creates a number of
defined to increase the solubility or dissolution rate challenges throughout drug discovery, beginning with
(realizing that the dependence of dissolution on primary activity screens and progressing through to
solubility dictates that increases in solubility inher- secondary in vitro assays and in vivo assessment.
ently increase the dissolution rate): In many in vitro assays, there is little scope for im-
proving solubility given the poor tolerability of many in
• Reducing intermolecular forces in the solid state
vitro biologic test systems for solubilizing components.
(increases solubility and dissolution rate)
In this instance, the focus must be on understanding
• Increasing the strength of solute-solvent interac-
the consequences of simple solution preparation and
tions in solution (increases solubility and dissolu-
manipulation (e.g., dilution), appreciating the poten-
tion rate)
tially confounding effects of compound precipitation on
• Increasing the surface area available for dissolu-
assay results, and grasping the impact of commonly
tion (increases dissolution rate, potential to mod-
used buffer components in dictating “free” compound
erately increase solubility at very small particle
concentrations in solution. The sections that follow
sizes (,1 mm)
highlight some of the common problems associated with
the in vitro testing of poorly water-soluble compounds
F. Summary and, where available, approaches that can be taken to
Low aqueous solubility and reduced dissolution overcome, or at least minimize, these issues.
rates are a common property of many new drug
candidates, and these properties create a number of A. Drug Precipitation, Adsorption, Binding, and
challenges during drug discovery and development. Complexation in In Vitro Assays
An understanding of the determinants of solubility Beyond the realms of chemical synthesis, most in
and dissolution provides a framework from which vitro evaluations of compound performance are con-
approaches to enhance solubilization may be devel- ducted using aqueous-based biologic buffers and re-
oped. In subsequent sections of this review, we first lated media. The range of in vitro testing protocols is,
address the complexities of working with poorly of course, broad but might include binding or displace-
water-soluble drugs in vitro and subsequently sum- ment assays, enzyme inhibition studies, activity screen-
marize the approaches that can be taken to assist in ing in cell culture, traditional organ-bath pharmacology,
the development of both parenteral and oral formula- assessment of uptake and transport in cell culture
tions. The main body of the review follows and systems, or excised tissues and metabolic stability
provides a detailed account of the technological studies using microsomes or hepatocytes. In all cases,
approaches that can be taken to support the de- an accurate knowledge of the concentration of drug in
velopment of effective formulations for poorly water- solution is required as it is critical to the determination
soluble drugs. Comment is made as to the many of the experimental endpoint. For example, most in
different approaches that might be taken during lead vitro assays of drug activity are based on a concentra-
optimization and preclinical development and also tion-response relationship, with the endpoint being
those strategies that are also appropriate for exten- some measure of the inhibitory or effective concentra-
sion into clinical development and ultimately to the tion (e.g., IC50 or EC50). In other assays, changes to drug
market. To constrain the scope of this review, synthetic concentrations in solution during the assay are used as
medicinal chemistry approaches to solubility manipula- an indicator of cellular uptake or transport, metabolism,
tion are not addressed and the discussion is limited to or binding, all of which rely on a known concentration of
approaches that do not result in the generation of compound in solution.
a fundamentally new chemical entity. For more in- In most drug discovery settings, moderate- to high-
formation on approaches to solubility manipulation via throughput assay formats (i.e., 96-, 384-, or 1536-well
structural modification, the interested reader is directed plates) are used, and compounds are introduced into
to the following reviews: Fleisher et al. (1996), Stella and aqueous biologic media via dilution of concentrated
Strategies for Low Drug Solubility 327

stock solutions prepared in water miscible organic bound, complexed, or precipitated drug is a common
solvents, the most common being dimethyl sulfoxide procedure during the conduct of many in vitro assays.
(DMSO). DMSO is an excellent solvent for many poorly Assessment of the potential for adsorption during
soluble compounds, including those with diverse filtration is an important validation step but is
chemical structures, and allows for the preparation of particularly critical for poorly water-soluble drugs.
highly concentrated stock solutions for subsequent Where significant adsorption is evident, and unavoid-
dilution for most compounds. However, compound able, then it may be necessary to avoid the process of
precipitation after dilution of concentrated cosolvent filtration. Equilibrium dialysis may provide an alter-
solutions is common (see Section VI) and can lead to native under such circumstances, but the potential for
variable responses depending on how the dilution is adsorption to the dialysis membrane is also high. A
conducted and the composition of the final assay final approach is to use ultracentrifugation to separate
media. This in turn can lead to a high degree of larger drug complexes from free drug in solution, but
variability and inconsistent results between assays these measurements are time consuming, require
where these variables may be different. Different specialized equipment and a significant density differ-
biologic assays also vary in their tolerance to the final ence between species, and ultimately still carry the risk
DMSO (or other cosolvent) concentrations, making it of drug adsorption to centrifuge tubes or plates.
difficult to standardize experimental conditions and Several different approaches can be taken to over-
dilution procedures across different assay formats. come issues of adsorption. The first is the choice of tube,
For many in vitro assays (with the exception of high- filter, culture flask, multiwell plate, filter or pipette tip,
throughput assays specifically designed to assess as many manufacturers supply materials with modified
solubility), it may be difficult, if not impossible, to surface properties to reduce nonspecific adsorption. It
detect the presence of finely precipitated material on is important to appreciate the consequences of using
dilution into aqueous media, and measuring the final different surfaces as those specifically designed to, for
concentration of drug may be impractical. Working example, promote cell adhesion by making them more
with compounds with inherently low aqueous solubilities hydrophobic, will also increase the likelihood of non-
generally limits the available range of drug concentra- specific drug adsorption. In contrast, more hydrophilic
tions that can be used to define concentration-dependent surfaces will generally reduce adsorption by reducing
processes. Under these circumstances, complete binding the thermodynamic favorability of drug leaving the
or inhibition profiles may be difficult to define since the largely aqueous solution environment. Another ap-
solubility limit is reached before saturation of binding proach involves pre-exposing potential adsorption
sites or approach to maximal inhibition. sites to drug solution with a view to saturating adsorption
Furthermore, the physicochemical properties that before conduct of the experiment. Finally, alteration of
predispose compounds to low aqueous solubility can the solution properties can reduce adsorption by in-
also lead to an association of drug molecule with creasing drug affinity for the bulk solution and reducing
hydrophobic environments and surfaces. In fact, this the effective partition coefficient between solution and
phenomenon is often a driving force for biologic potency drug adsorption sites. Most commonly, this is achieved
since partitioning into and across cell membranes and via the inclusion of small quantities of a cosolvent
interaction with cellular and molecular targets are (depending on the sensitivity of the particular assay) or
thermodynamically favored compared with residence by manipulating solution pH to increase solute-solvent
within an aqueous solution. However, these character- interactions in solution. The use of pH and cosolvents to
istics also lead to an inherent propensity for non- enhance solubility is described in more detail in
specific adsorption to surfaces such as tubes, pipette Sections IV and VI, respectively.
tips, filters, syringes, multiwell plates and cellular Adsorption may also be reduced via the addition of
supports. Under these circumstances, a decrease in a complexation or solubilization agent such as a sur-
drug concentration in solution may be reflective of factant (see Section VII), cyclodextrin (see Section
nonspecific adsorption rather than binding, uptake, or VIII), or protein. Although these approaches are often
transport, artificially reducing the concentration of highly effective, they introduce a further complexity,
drug in solution and leading to an erroneous endpoint namely, changes to the chemical potential or thermo-
determination if concentrations are assumed on the dynamic activity of free (unbound) drug in solution as
basis of only the dilution factor. Clearly, there are discussed in the following section.
advantages to obtaining measured concentrations to
provide confidence in the experimental results when B. Changes to Thermodynamic Activity Resulting
practical and also to provide an assurance of mass from Complexation, Binding, or Solubilization
balance, which is a critical control measurement in any The effective concentration of a species in solution is
mass transport experiment. most accurately defined by its thermodynamic activity
The adsorption of drug to filter membranes requires (a), which is related to the concentration (C) and the
special mention since separation of free drug from activity coefficient (g):
328 Williams et al.

a 5 g×C ð11Þ In addition to the use of surfactants, other common


circumstances where the thermodynamic activity of
A detailed evaluation of thermodynamic activity and a compound might be reduced include the introduction
chemical potential is beyond the scope of this review, of a complexation agent, such as a cyclodextrin, or the
but in simple terms, the effective concentration or addition of plasma or serum proteins to cell-based
thermodynamic activity can be considered as the con- assay media. For the latter situation, the so-called
centration of drug in solution that is unconstrained by serum shift phenomenon is widely recognized where in
interaction with any other molecular species and is vitro activity changes in response to changing concen-
therefore available to exert its effect, regardless of trations of serum present in the media. For a more
whether the effect is to bind to a receptor or to diffuse comprehensive review of the effect of protein binding
across a membrane. In concentrated solutions, for on the pharmacological activity of drugs and the com-
example, the close molecular proximity of drug molecules plexities of interpreting static (in vitro) versus dynamic
to each other may promote solute-solute interactions (in vivo) situations, the reader is referred to the
(i.e., enhance intramolecular association), thereby re- excellent review by Smith et al. (2010). In each of
ducing thermodynamic activity. Under these circum- these cases, the total concentration of drug is signifi-
stances, the activity coefficient is less than unity, and cantly higher than the effective or free concentration,
the effective concentration or activity is less than the and the situation is further confounded by the difficulty
measured concentration. In typical dilute solutions, the in accurately quantifying free drug concentrations. In
degree of dilution is expected to reduce solute-solute contrast, solubility enhancement through pH manipula-
intermolecular interactions in solution such that each tion or cosolvency will typically have limited impact on
molecule effectively acts independently, and under thermodynamic activity of drug in solution.
these circumstances, the activity coefficient is unity
and activity is equal to concentration. This allows the
definition of equilibrium constants, for example, in III. In Vivo Assessment of Poorly
dilute solution using drug concentration instead of the Water-Soluble Compounds
thermodynamically correct use of activities. In vivo assessment of new drug candidates begins
The relevance of this discussion to the in vitro with early in vivo efficacy studies in animal models and
testing of poorly water-soluble drugs is that many continues through pharmacokinetic and dose-limiting
strategies that promote drug solubility in aqueous toxicology studies in various preclinical species and
solution and assay media can change the thermody- ultimately into human clinical trials. Preclinical
namic activity of the drug in solution. This is most pharmacokinetic studies are heavily used to support
readily illustrated by considering the impact of the human dose and pharmacokinetic predictions and,
addition of a surfactant to an aqueous solution of drug. along with results from toxicology studies, are used to
Surfactants are amphiphilic molecules, and in aqueous select a safe starting dose in humans. For in vivo
solutions, they can exist as either monomers or as evaluations in animals, there are multiple approaches
micellar structures. As surfactant concentrations are that can be used to facilitate i.v. administration and to
increased above the critical micelle concentration improve exposure after oral dosing by the use of
(CMC), the concentration of monomeric surfactant enabling formulations. However, there is always a risk
remains constant while the concentration of surfactant that early incorporation of an enabling formulation
present as micelles in solution increases. Surfactant approach may shift the focus away from structural
micelles typically enhance drug solubility by providing optimization. Even in circumstances where lead opti-
a hydrophobic environment in the micellar core to mization strategies are successful in identifying drug
solubilize poorly water-soluble drugs as they have candidates with improved solubility properties, it is
a greater affinity for this hydrophobic environment than likely that at some stage during drug discovery there
for the surrounding aqueous environment (see Section will be the need to assess less soluble early leads for
VII). However, under this circumstance, the thermody- intrinsic activity and proof of concept efficacy to justify
namic activity of the drug is much lower than the total compound or series progression.
concentration as the activity coefficient of the drug is A complication of poor aqueous solubility in the in
lower than unity. In the case of drug solubilized in vivo assessment of compounds throughout drug dis-
a surfactant micelle, drug can be considered as existing covery and development is that compound supply is
in equilibrium between solubilized drug within the frequently limited (particularly in early discovery), and
micelle and free drug in an intermicellar phase or bulk material that is available most likely will not have the
solution phase. Here, solubilized drug is highly con- final, or optimal, solid-state properties. As compounds
strained by the surrounding micellar structure, and the progress through discovery and into development, the
effective concentration (i.e., the thermodynamic activity) solid-state properties will almost certainly change;
is more accurately represented by the free (nonmicellar) crystal forms will become better defined, particle size
concentration of drug. reduction may be introduced, and salt forms will likely
Strategies for Low Drug Solubility 329

become available for compounds with ionizable func- soluble drugs, if possible, because of the potential for
tional groups. The timing at which these changes occur precipitation. Addition of small quantities of surfactants
during development will impact early preclinical data, (,0.5%) or polymers may assist in preventing pre-
and this needs to be appreciated and factored into cipitation at the injection site.
study design and data interpretation. Drug administration via other parenteral routes,
including s.c., i.m., and i.p. administration, is on
A. Parenteral Administration occasion warranted, particularly in proof-of-concept
1. Complexities with Parenteral Administration of biology studies where there is evidence of poor oral
Poorly Water-Soluble Drugs. In comparison with the exposure resulting from solubility limitations. With
large number of drugs intended for oral administration, non-i.v. parenteral routes, the potential for the route
parenteral formulations constitute a more limited pro- and the formulation vehicle to have an effect (either
portion of marketed products. However, the generation desirable or undesirable) on the absorption profile is
of useful parenteral formulations remains a key compo- greater, and depot effects are often seen. For these
nent of drug discovery and drug development for almost administration routes, solution formulations are still
all drug molecules (regardless of the intended route of preferred since the volume of fluid at the injection site
clinical administration) since data obtained after i.v. is low, and, therefore, dissolution of suspension for-
administration is necessary to generate fundamental mulations may be limited. The lack of formulation
pharmacokinetic parameters (volume of distribution, dilution at the injection site dictates that the range of
clearance, half-life, absolute bioavailability) to support excipients and conditions to promote drug solubiliza-
compound optimization and progression, and such data tion is even more limited than it is after i.v. admini-
are also useful to support early stage pharmacodynamic stration. Ideally, non-i.v. parenteral formulations
assessment of drugs intended for acute administration. should be as close as possible to physiologic pH, and
Low aqueous solubility significantly complicates the concentrations of cosolvents should be minimized.
development of i.v. formulations since, in almost all 2. Parenteral Formulation Approaches for Poorly
cases, simple solution formulations are required. The Water-Soluble Drugs. The most common approaches
following section describes several different approaches for the development of parenteral formulations for
that can be used as i.v. solution formulations, depending poorly water-soluble drugs are summarized in the
on the physicochemical properties of the drug and the following discussion, and strategies for rapid identifi-
intended use (e.g., animal or humans). Depending on cation of the most appropriate approach are discussed.
the formulation strategy adopted, increases in solubility The acceptable limits for each of these will depend on
of several orders of magnitude are possible. The risks many factors including the species (animal or human),
associated with i.v. administration of poorly water- dose volume, rate of administration (e.g., infusion or
soluble drugs include, first, the potential for drug bolus), parenteral route (i.v., i.m., s.c., i.p.), and duration
precipitation on administration and, second, the poten- of treatment (Gad et al., 2006; Li and Zhao, 2007). A
tial for formulation components to cause pain, phlebitis, summary of parenteral formulation approaches is given
inflammation, hemolysis, or unacceptable toxicology. in Table 2, and these are briefly described as follows.
Precipitation is more likely for formulations where pH pH adjustment (see Section IV) is a powerful
adjustment or cosolvents form the basis for drug solu- mechanism by which the solubility of weak acids and
bilization since both are likely to be altered significantly weak bases may be enhanced, but the approach is
on dilution (see Sections IV and VI). In contrast, sur- limited by the pKa of the compound and the acceptable
factant solubilization, complexation, and i.v. emulsions pH range of the formulation. In this regard, acidic
are far less sensitive to dilution (see Sections VII, VIII, solutions are generally more readily tolerated than are
and XI). Where dilution-mediated precipitation is basic solutions, and working pH ranges of 2–9 (Li and
possible, slow administration into large veins, where Zhao, 2007) or 4–9 (Lee et al., 2003) have been
the blood flow is higher (and therefore the supply of suggested. The pH of parenteral solutions can be
plasma proteins and lipoproteins to provide in vivo manipulated via the addition of small quantities of
binding is higher) is preferred, and this can also reduce strong acids or bases, such as HCl or NaOH but is more
the possibility of pain on injection, hemolysis, and usefully achieved via the use of a buffer, generally with
phlebitis since the irritant component is maximally as low a buffer concentration as possible while still
diluted. For studies in rats and dogs, this can be maintaining the desired pH. Common buffer systems
accomplished by implanting a dosing cannula into the used in parenteral formulations include phosphate and
jugular vein (or another large vein) with administration bicarbonate at neutral or slightly basic pH and citrate,
over a period of 5 to 10 min for rats and up to 30 min or lactate, tartrate, acetate, and maleate for more acidic
longer in dogs. In contrast, administration into small solutions (Flynn, 1980; Kipp, 2007).
veins in the extremities (where blood flow is much Cosolvents (see Section VI) are also commonly used
slower), such as the tail vein in rodent models, and rapid to promote solubility in parenteral formulations, and
bolus injection, should be avoided with poorly water- ethanol, propylene glycol (PG), low-molecular-weight
330 Williams et al.

TABLE 2
Potential hierarchy of formulation approaches for the development of parenteral formulations for poorly water-soluble drugs
Previously published suggestions for maximum allowable quantities provided in parentheses.a,b,c Further details are provided in individual sections. Note that common
examples are provided, and this is not intended to represent an exhaustive list.

Class Examples

Simple isotonic solutions Normal saline


5% dextrose
pH-adjusted solutions pH 6–9 for weak acids; common buffers include phosphate, bicarbonate
pH 2–6 for weak bases; common buffers include citrate, acetate, lactate
Cosolvents Examples:
Propylene glycol (30–60%)a (50%  0.5 ml/kg)b
PEG 400 (40–100%)a (50%  1 ml/kg)b
Dimethylacetamide (10–30%)a
N-Methylpyrrolidone (10–20%)a
DMSO (10–20%)a (100%  0.1 ml/kg)b
Ethanol (10%)a (20%  0.5 ml/kg)b
Common cosolvent combinations:
Propylene glycol:ethanol:water 40:10:50
PEG 400:ethanol:water 40:10:50
Propylene glycol:ethanol:dimethylacetamide:water 20:10:10:60
Cosolvents 1 pH adjustment As above plus buffered aqueous component
Surfactants Examples:
Cremophor EL (5–10%a (10%  0.5 ml/kg, single dose only, potential for anaphylaxis in dogs)b
Tween 80 (5–10%)a (2%  0.25 ml/kg, potential for anaphylaxis in dogs)b
Pluronic F68 (20–50%)a (15%  0.5 ml/kg)b
Complexation agents Hydroxypropyl-b-cyclodextrin (20–40%)a
Sulfobutylether-b-cyclodextrin (Captisol) (20–40%)a (20%  5 ml/kg)b
Surfactants or complexation agents As above plus buffered aqueous component
1 pH adjustment
Lipid emulsions Phospholipid stabilized soybean oil emulsion (10–20% lipid) (Intralipid)
Nanomilled drug crystals Nanocrystals
Advanced colloids Liposomes, polymeric or lipid nanoparticles
a
Concentration range based on rat or mouse and single administration at 10 ml/kg (Li and Zhao, 2007).
b
Concentration range and maximum dose volume for acute and subchronic doses in rodents (#7 days) (Neervannan, 2006).
c
Descriptions of toxicological outcomes after specific dose regimens in several species for many common excipients also available in Gad et al. (2006).

polyethylene glycol (PEG 300 or 400), dimethylaceta- TPGS, and poloxamers (Pluronics, including Pluronic
mide (DMA), and DMSO have been used. Combina- F68; BASF Corporation). Whereas surfactants provide
tions of cosolvents provide benefits since cosolvency is for relatively robust solubilization, a complexity is the
additive, whereas toxicity is often dependent on the growing realization that many surfactants can affect
particular solvent. transport processes, such as cellular efflux, and may
In many cases, judicious manipulation of solution pH also result in significant adverse effects, including
in combination with cosolvents is sufficient to provide hypersensitivity reactions (Gelderblom et al., 2001; ten
solubilization and is typically viewed as the first-line Tije et al., 2003). Cremophor EL, in particular, seems to
approach. For example, in a review of 317 discovery promote an immune reaction in some species (most
compounds, Lee et al., (2003) reported that .90% of notably dogs) when administered parenterally. Surfactants
compounds could be formulated using pH and cosol- bring additional complexities to data interpretation and,
vents and that .60% of these could be formulated where possible, might usefully be avoided in favor of pH
using cosolvent levels of less than 50% v/v. changes, cosolvents, or the use of cyclodextrins.
However, where pH adjustment and cosolvents fail Complexation agents (see Section VIII), such as
to provide the required solubility, alternative approaches cyclodextrins, provide a mechanism for enhancing
must be used. Second-tier approaches include the use of solubility for many drugs and have been increasingly
cyclodextrins, surfactants, and lipid emulsions and, used as modified cyclodextrins, such as hydroxypropyl-
ultimately, more complex formulations such as lip- b-cyclodextrin (HP-b-CD) and sulfobutylether-b-cyclo-
osomes, microemulsions, or nanosuspensions. For dextrin (SBE-b-CD), have become more cost effective.
more complex systems, however, the potential effect, HP-b-CD and SBE-b-CD have higher aqueous solubil-
either positive or negative, on systemic clearance and ities than do unmodified cyclodextrins (therefore allow-
distribution (and also absorption for non-i.v. paren- ing the support of higher concentrations of complexed
teral formulations) should be carefully considered. drug) and, importantly, are becoming well characterized
Surfactants (see Section VII) can also be used to in terms of their safety profile. In rare situations where
enhance drug solubility via micellar solubilization. the binding constant between drug and cyclodextrin is
Surfactants are generally limited to those that are particularly high (e.g., ;106 M21), there is the potential
nonionic because of their more favorable safety profile for alteration of the pharmacokinetic distribution and
and may include polysorbates (e.g., Tween 80), Cremo- elimination processes (Charman et al., 2006), but most
phors (including Cremophor EL and RH40), Solutol drugs exhibit significantly lower binding constants
HS-15 (BASF Corporation, Washington, DC), vitamin E (,104 M21) and are efficiently “released” from the
Strategies for Low Drug Solubility 331

complex via dilution and displacement by endogenous B. Oral Administration


species (Stella and Rajewski, 1997; Stella et al., 1999; In contrast to parenteral formulations, the ap-
Stella and He, 2008). HP-b-CD and SBE-b-CD are proaches that are commonly used to support the oral
commonly used in parenteral formulations in concen- delivery of poorly water-soluble drugs vary signifi-
tration ranges of up to 20 to 40% w/v in both humans cantly depending on the purpose of the in vivo study,
and animals, and combining complexation with pH the species in which studies are conducted, the quantity
adjustment can, in some cases, provide additional of material available, and the level of exposure required.
solubility enhancement. Distinction between the approaches taken at different
Intravenous emulsion formulations (see Section XI ) stages of the discovery-development continuum is dif-
can be used for lipophilic, poorly water-soluble drugs ficult, and different organizations will have different ap-
that have high solubility in lipids. Typically, lipid proaches depending on in-house experience and
emulsion formulations contain between 10 and 20% expertise. Nonetheless, the formulation strategies that
lipid, which limits drug load for compounds that do not are commonly adopted during discovery and develop-
have high lipid solubility. Intravenous emulsion for- ment are summarized as follows.
mulations have been used for many years as nutri- 1. Formulations to Support Drug Discovery. In
tional supplements (e.g., Intralipid; Baxter Healthcare, proof-of-concept pharmacology studies, the focus is
Deerfield, IL) and are usually formulated using soybean generally placed on generating sufficient exposure to
oil as the dispersed phase and phospholipids as the allow early indication of the efficacy of a particular
emulsifier. Emulsions can be formulated de novo by pharmacophore or compound series. At this stage, only
dissolving drug into the oil phase of the emulsion before small quantities of compound are usually available,
emulsification; however, this requires the use of and control of crystallinity is limited, with early batches
specialized high-pressure homogenization equipment often including a significant proportion of amorphous
to provide emulsions with submicron particle sizes. material. Before a salt screening program, ionizable
Alternatively, drug may be incorporated into preformed compounds may also be available only as the free acid or
commercial emulsions by dissolving the drug in a small base or potentially as a non-optimal salt form. Under
volume of cosolvent with subsequent slow incorporation these circumstances, simple solution formulations are
using sonication (El-Sayed and Repta, 1983). preferred since they circumvent the complexities of
More complex formulation approaches, such as the dissolution and, therefore, obviate any variability in
use of nanosuspensions (see Section IX), liposomes, or exposure resulting from differences in solid-state
microemulsions, have all been used to facilitate characteristics. Solution formulations facilitate ease
parenteral administration but can substantially affect and reproducibility of administration via oral gavage,
drug disposition. These approaches are better avoided enable ready conduct of single and multiple dose
when the intent of the formulation is to define studies, and support dosing strategies such as cassette
fundamental pharmacokinetic parameters of a drug. dosing. The approaches taken to develop oral solution
In many cases, combination approaches to solubili- formations are similar to those used for parenteral
zation in parenteral formulations are beneficial. The formulations, however, the acceptable limits on formu-
overarching premise for this approach is that the lation components are typically wider. A useful order of
benefits of solubilization are often additive, whereas potential formulation approaches to achieve simple
the effects on toxicity are typically a function of reach- oral solution formulations is provided in Table 4. In early
ing a critical level for a single component. Maintaining stage proof-of-concept studies, relatively aggressive ab-
concentration of one solubilizer below that threshold, sorption enabling formulations may be used to overcome
but combining multiple approaches, therefore, may solubility and dissolution rate limitations, depending on
have benefit. In some cases, however, combination ap- the frequency and length of dosing; however, caution
proaches add complexity without generating benefit or, should be used to ensure that the formulation components
indeed, are detrimental. Combination approaches are do not confound the efficacy readout, and vehicle controls
summarized in Table 3 and are further described in are essential.
detail in each individual section of this review. Oral delivery during lead optimization is typically
focused on prioritizing lead compounds from within
TABLE 3
Summary effect of combining two common solubilization approaches for a series with more favorable absorption properties and
parenteral formulations on identifying the key structural and physicochemical
pH Adjustment Cosolvents Surfactants Cyclodextrins factors that contribute to poor oral bioavailability, in
pH adjustment ♢♢♢ ♢♢♢ ♢♢♢ an effort to guide subsequent structural modification.
Cosolvents ♢♢♢ ♦♦♦ ♢♢♢ Understanding the basis for poor oral absorption is
Surfactants ♢♢♢ ♦♦♦ ddd
Cyclodextrins ♢♢♢ ♢♢♢ ddd fundamental to designing a strategy to overcome this
♢ ♢ ♢, additive effect on drug solubilization; ♦♦♦, no clear effect on drug
limitation. The approaches used for oral formulations
solubilization; ddd, possible decrease in drug solubilization. of poorly water-soluble compounds in lead optimization
332 Williams et al.

TABLE 4
Possible hierarchy of formulation approaches for the development of oral solution formulations
Previously published suggestions for maximum allowable quantities of formulation excipients provided in parentheses.a,b,c Exemplar formulations—for example, cosolvent
combinations, surfactant formulations, and lipid-based formulations—are provided, realizing that they represent a limited “snapshot” of possible formulations, concentration
ranges, etc. (Further details are provided in the sections of this review that describe each technology.)

Class Examples

Simple Solutions 6 pH As Table 2


Cosolvents 6 pH Propylene glycol (30–60%a (80%  2 ml/kg)b
PEG 400 (40–100%)a (100%  2 ml/kg)b
DMA (10–30%)a
N-Methylpyrrolidone (10–20%)a
DMSO (10–20%)a (50%  0.5 ml/kg]b
Ethanol [10%]a [50%  0.5 ml/kg]b
Oral cosolvent combinations are common e.g., 75% PEG 400: 25% propylene glycol
Surfactants As Table 2, but also solid and semisolid surfactants including:
a
D-a-Tocopheryl polyethylene glycol 1000 succinate (vitamin E TPGS 1000) (20–50%)
Lauroyl macrogol-32 glycerides (Gelucire 44/14) [20–50%]a
Caprylocaproyl macrogol-8-glycerides (Labrasol) [40–60%]a
Complexation agents As Table 2
Cosolvent/surfactant PEG-based cosolvent/surfactant combinations, e.g., 55% PEG: 25% propylene glycol: 25% Cremophor EL
combinations Dispersed surfactant/cosolvent combinations, e.g., 50% Cremophor EL: 25% ethanol: 25% propylene glycol 1:1
in water or 20% PEG 400: 10% Cremophor EL: 10% ethanol in water
Simple lipid formulations Lipid solution formulations, e.g., 100% soybean oil, corn oil
Self-emulsifying products, e.g., Labrasol, Labrafil, or Gelucire up to 50% in water
Self-emulsifying preparations, e.g.;

30% long-chain triglyceride (e.g., soybean oil): 30% long-chain diglycerides/monoglycerides


(e.g., Maisine 35-1): 30% Cremophor EL: 10% ethanol, usually dispersed in water, or,

30% medium-chain triglyceride (e.g., Captex): 30% medium-chain diglycerides/monoglycerides


(e.g., Capmul): 30% Cremophor EL: 10% ethanol, usually dispersed in water
a
Concentration range based on rat or mouse and single administration at 10 ml/kg (Li and Zhao, 2007).
b
Concentration range and maximum dose volume for acute and subchronic doses in rodents (#7 days) (Neervannan, 2006).
c
Descriptions of toxicological outcomes after specific dose regimens in several species for many common excipients are also available in Gad et al. (2006).

are therefore usually less complex than those incor- use of suspension formulations is the potential impact of
porated, for example, during discovery biology, and are particle size on dissolution rate and the possibility of
designed to discriminate between compounds in the batch-to-batch variability given that solid-state proper-
absence of enabling formulation effects. Consistency in ties and particle size are not likely to be optimized at the
formulation is important to distinguish between poten- time the initial pharmacokinetic and bioavailability
tial effects resulting from the formulation and intrinsic studies are conducted. Assessment and monitoring of
differences in structural and physicochemical properties particle size are beneficial, and normalization of particle
of the lead. size across batches via laboratory scale milling may
Simple suspension formulations containing a sus- provide greater consistency.
pending agent (e.g., cellulose derivatives) along with Another approach is to use simple solution formula-
a low concentration of surfactant to reduce particle tions, typically prepared using cosolvents with or with-
agglomeration are often used in early bioavailability out pH manipulation. These formulations have the
studies (see Table 5). These formulations provide little advantage of being independent of solid-state proper-
in the way of “assisted” dissolution and therefore tend ties since compounds are in solution but have the
to discriminate between different compounds. A fur- disadvantage of likely precipitation after administra-
ther advantage is that they can be readily extended tion, particularly if the formulations prepared using
into animal toxicology studies, provided sufficient com- cosolvents are near the saturation solubility limit.
pound exposure can be achieved. A complexity in the Generally, more specialized enabling formulations, such
as surfactant-, cyclodextrin- and lipid-based formula-
TABLE 5 tions, are not ideal for lead optimization in that they
Typical suspension formulations may mask intrinsically poor physicochemical properties
Class Examples and take the focus away from structural optimization to
Simple suspensions 0.5–2% cellulose-derived suspending agents improve solubility. An early reliance on the use of
(e.g., methylcellulose, hydroxypropyl enabling formulations also complicates subsequent de-
methylcellulose, hydroxypropyl cellulose,
or carboxymethyl cellulose) plus
velopmental activities.
Up to 10% surfactant (e.g., Tween 80, As larger quantities of drug become available, more
Pluronic F68) definitive information regarding solid-state properties
With or without pH adjustment
Milled suspensions As above with particle size reduction (milling) is generated and salt screening programs are initiated
Nanosuspensions Surfactant and/or polymer stabilized for weak acids and bases (see Section IV). Isolation as
nanosuspension (e.g., NanoCrystal)
a salt form can provide for significant enhancements in
Strategies for Low Drug Solubility 333

dissolution rate. The availability of larger quantities of a relatively small scale, hand filling of dry powder
material also allows more complex particle size re- blends or granules into hard gelatin capsules or liquid
duction strategies to be explored including micronized filling into preformed soft gelatin capsules can be used.
formulations and nanosuspensions (see Section IX). Sealed hard-gelatin capsules provide an alternative
2. Use of Enabling Formulations to Promote Oral for liquid fills and may be sealed by hand or, more
Absorption. In situations where attempts to increase effectively using specialized capsule-sealing equip-
polarity and water solubility lead to significant de- ment, realizing that such equipment may not be readily
creases in potency, and where solubility or dissolution available at this stage. For poorly water-soluble drugs,
has been identified as the rate-limiting process for the shift to encapsulated and solid dosage forms dictates
absorption, a strategic decision may be made to progress renewed concentration on solid-state formulation strat-
a compound series via the use of enabling formulations egies such as salts, cocrystals, and particle size reduc-
to improve bioavailability and exposure. This approach tion alongside more complex formulation approaches,
may be taken in an effort to obtain rapid proof-of- including solid dispersions and lipid-based formulations.
concept data in humans against a new target or simply 3. Preclinical Toxicology Formulations. Preclinical
to enable early pharmacokinetic and safety data to be toxicology evaluation is a particular challenge for poorly
obtained for a novel compound series. water-soluble drugs. The need to escalate doses to the
The preclinical assessment of more specialized point where in vivo exposure is sufficient to demon-
oral formulations will depend heavily on the type of strate limiting toxicity generally requires absorbed
formulation being assessed. Generally, rodents require doses (in milligram per kilogram terms) 30 to 100 times
liquid formulations that can be easily administered via greater than might be envisaged with a clinically
a gavage tube. Recent developments in the availability relevant dose and formulation. Where water solubility
of minicapsules and tablets allows dosing of tablets or is low, this is often impossible with solution or suspen-
encapsulated materials to rodents; however, the re- sion formulations, since beyond a certain point, drug
quirement for specialized equipment to prepare these exposure remains constant regardless of increases in
dosage forms and the low liquid volumes in the rodent dose, dictating the need for more complex formulation
GI tract (which may limit the rate of capsule/tablet approaches; however, this requirement is tempered by
disintegration) make this approach less common. potential concerns regarding the toxicology of the large
Alternate crystal forms such as cocrystals, polymorphs, quantities of excipient often required to afford the
or amorphous drug may explored, although the inherent higher exposure. Relatively little information is publi-
instability of metastable crystal and amorphous forms cally available that describes the quantities of different
dictates the need for close monitoring of crystallization combinations of excipients that do not lead to toxic
properties, and formulation approaches to stabilize endpoints over a range of dosing periods. One notable
against solid-state transitions may be necessary, depend- exception is an excellent review by Gad et al. (2006),
ing on the required stability period. Cocrystals provide which provides a summary of the effects of a range
some advantage in this respect as their crystal lattice of excipients in toxicity testing protocols of different
energy is significantly higher than amorphous drug, lengths. Neervannan (2006) also provide some data
allowing the generation of formulations with improved regarding the maximum quantities of common exci-
stability. Stabilization of polymorphs or amorphous drug pients that might be safely used in acute and subchronic
may be achieved by the use of (usually polymeric) solid dosing protocols for up to 7 days. However, these
dispersion formulations, where reduced molecular mo- communications provide data on a limited number of
bility, often achieved via intermolecular interactions surfactants, lipids, and polymers and do not address the
between carrier and drug, reduces the risk of changes potential toxicity of the combinations of components
to the amorphous-crystal structure during storage. Solid often used.
dispersions require relatively complex equipment to Toxicity concerns for many of the excipients used in
provide clinical formulations, such as spray drying or more complex formulations are attenuated by the
hot melt extrusion; however, trial batches can be realization that many polymers and surfactants have
formulated using simple benchtop equipment such as high molecular weights and are relatively polar, or
a rotary evaporator or freeze-driers. amphiphilic. As such, the extent of absorption of these
As drug candidates progress through development, excipients after oral administration is low. Many sur-
preclinical investigations in nonrodent species will be factants are also digestible (Cuiné et al., 2008). Under
conducted, usually in dogs or nonhuman primates. The these circumstances, the risks of systemic toxicity are
physical form of materials used in these studies will low, and the oral LD50s of most hydrophilic polymers
vary, and simple solutions or suspensions will often and nonionic surfactants are high (Rowe et al., 2009).
still be used. Larger species generally provide the first The major concerns therefore revolve around the
opportunity to assess the performance of prototype possibility of GI side effects, including loose stools and
tablet and capsule formulations that might ultimately emesis (Reppas et al., 1991; Andersen, 1994; Schulze
progress to the clinic. For ease of preparation on et al., 2005; Ruble et al., 2006). By incorporation of
334 Williams et al.

vehicle controls, many of these effects can be accounted TABLE 6


Formulation approaches for the development of solid oral formulations
for during data analysis. Confounding data interpreta-
tions issues arise, however, where GI symptoms man- Class Examples

ifest as a consequence of the candidate drug’s desired Traditional tablets, Isolation of salt forms or cocrystals to enhance
capsule fills solubility
biologic activity or toxicity profile. Addition of small quantities of solid surfactant
Cosolvents, in particular ethanol, typically exhibit to enhance wetting
more significant toxicity profiles and, hence, cosolvent Particle size Milling or nanosizing before encapsulation
reduction or tableting
combinations are commonly used to maintain the Lipid-based Liquid fill or thermo-softening fills in soft
concentration of one single component at a relatively formulations gelatin capsules or sealed hard gelatin
capsules
low level. Finally, several surfactants have also been Solid dispersions Usually drug-polymer combinations that may
shown to inhibit efflux transporters (Batrakova et al., be generated by spray-drying or melt-
extrusion, commonly used to stabilize
1999; Yamagata et al., 2007a) and GI cytochrome P450 amorphous drug
enzymes (Wandel et al., 2003; Bravo González et al.,
2004; Christiansen et al., 2011) and may therefore
affect drug bioavailability after oral administration.
Evidence of systemic effects, however, is limited except salt formation, cocrystals, or nanocrystals fail to pro-
after parenteral administration. After parenteral admin- vide sufficient exposure, the major routes of progres-
istration, surfactants such as Tween 80 and Cremophor sion are the use of lipid-based formulations or solid
EL may stimulate hypersensitivity reactions (Weiss dispersions. The choice of approach is dictated largely
et al., 1990; Gelderblom et al., 2001; ten Tije et al., by the physicochemical properties of the drug candidate,
2003), especially in dogs (Lorenz et al., 1982). where lipid-based formulations lend themselves to
4. Development of Clinical Formulations for Poorly poorly water-soluble, lipophilic drugs (i.e., those with
Water-Soluble Drugs. The strategies used when de- sufficient solubility in lipids, surfactants, and cosolvents
veloping clinical formulations typically fall into two to allow the target dose to be dissolved in the capsule fill
categories. The first strategy is the development of material), whereas solid dispersions are commonly
simple formulations that maximize the speed to first- explored for drug candidates with insufficient lipophi-
in-human studies but provide no functional link to the licity to allow the use of lipid-based formulations.
likely formulations used in later stage clinical trials In conclusion, the general strategies that may be
and that might be intended for commercialization. taken to identify and overcome the challenges of low
Second is the development of mature formulations, water solubility in the drug discovery and development
with the expectation of progressing essentially similar environment are summarized in the proceeding dis-
formulations through clinical development to commer- cussion. The tools that have been suggested as approaches
cialization. Where possible, the former approach uses to overcome the limitations of poor solubility are described
nonformulated drug to simplify dosage form develop- in detail in following sections.
ment, and, where solubility permits, simple direct
filling of drug in capsules or drug-in-bottle approaches
IV. Buffers and Salt Formation
can be used. The relatively recent application of direct
capsule-filling instruments such as Xcelodose (Capsu- Adjusting solution pH, usually via the use of a buffer,
gel, Greenwood, SC) that facilitate accurate filling of provides an effective means of increasing the proportion
doses ranging from 100 mg to 100 mg have greatly of a weakly acidic or basic drug that is present in the
simplified this process. ionized form. In turn, increasing ionization increases
For many poorly water-soluble compounds, however, polarity and therefore increases drug solubility in polar
low solubility limits the possibility of adopting an (aqueous) solutions. Salt formation results in a similar
accelerated route to first-in-human studies, and in endpoint, and in many ways salts of weak acids and
these instances, more sophisticated formulations are bases are the solid-state equivalent of pH adjustment.
required to ensure exposure. Formulations with the Salts are formed via an ionic interaction between
potential to progress through clinical trials, thereby weakly acidic or basic drugs and an oppositely charged
circumventing the need for repeated bioequivalence basic or acidic counterion. When salts are subsequently
bridging studies, are typically evaluated. The escala- allowed to dissolve and dissociate in water, the acidic or
tion of formulation complexity in this case (Table 6) basic counterion from which they were derived causes
commonly follows a pathway similar to that previously a shift in the pH of the solution to provide an
described for preclinical formulations, although in this analogous endpoint to that obtained by pH adjust-
case, the potential to progress through to full scale ment. Isolation of drugs as a particular salt form also
tableting or encapsulation methods and the use of changes the nature of the crystal lattice, resulting in
Good Manufacturing Practice processing equipment differences in dissolution rate from alterations in
becomes more important. Where simple solid-state or solid-state properties compared with the free acid or
particle size approaches to enhanced solubility such as free base (Huang and Tong, 2004; Corrigan, 2007).
Strategies for Low Drug Solubility 335

Diclofenac provides a good example of the potential base (BH1) (right equation) in water, with the
advantages of salt formation. In the un-ionized form, associated expressions for Ka shown in eq. 13.
diclofenac has a low aqueous solubility (,0.02 mg/ml)
Ka Ka
(Fini et al., 1999), whereas the sodium or potassium salt HA 1 H2 O ⇌ H3 O 1 1 A 2 BH 1 1 H2 O ⇌ H3 O 1 1 B
forms of diclofenac are .400 times more water soluble ð12Þ
(Fini et al., 1996). Similar trends are also evident with
weak bases (which make up the larger proportion of ½A 2 ½H3 O 1  ½H3 O 1 ½B
ionizable drug candidates); for example, the aqueous Ka5 Ka5 ð13Þ
½HA ½BH 1 
solubility of enalapril maleate is 25 mg/ml (Kasim et al.,
2004), whereas that of the free base is ;0.21 mg/ml. In As the dissociation constant may vary by several
unbuffered solutions, salts therefore provide an oppor- orders of magnitude, the pKa or negative logarithm of
tunity for significant increases in dissolution and Ka (pKa 5 2log Ka) is more typically used to compare
solubility. In contrast, in buffered solutions that favor acids and bases (Table 7). Decreasing pKa denotes
the un-ionized free acid or free base, the aqueous stronger acids (i.e., species that more readily donate
solubility advantages of pharmaceutical salts are a proton), whereas increasing pKa reflects stronger
attenuated, although significant increases in dis- bases (i.e., species that more readily accept a proton).
solution rate are still evident since the pH in the The Henderson-Hasselbalch equation (eq. 14) further
unstirred water layer reflects the pH of the dissolving describes the pH dependency of the ionization properties
salt rather than the bulk (buffered) pH (Serajuddin of weak acids (left equation) or bases (right equation) and
and Jarowski, 1985b; Li et al., 2005b; Serajuddin, relates the degree of ionization to the pH of a solution and
2007). the pKa of the drug.
Exploiting the enhanced solubility and dissolution
 2   
rate of the ionized form of a drug, either through pH ½A  ½B
pH 5 pKa 1 log pH 5 pKa 1 log
adjustment or salt formation, is therefore a common ½HA ½BH 1 
approach for oral and parenteral routes of drug
ð14Þ
delivery and should be regarded as a primary strategy
for addressing low aqueous solubility (Berge et al., From eq. 14, it is apparent that under conditions
1977; Stahl and Wermuth, 2002; Tao et al., 2009). where the pH of a solution is the same as the pKa of
Although this approach is applicable only to ionizable a dissolved weak acid or base, the concentrations of
compounds, weak acids and weak bases account for up ionized and un-ionized species are equivalent. A change
to two-thirds of marketed drugs and drugs in de- in pH by one log unit away from the pKa subsequently
velopment (Stahl and Wermuth, 2002); consequently, results in a 10-fold change in the degree of ionization.
the strategies described here are relevant across a wide 2. pH Solubility Relationships for Weak Electrolytes
range of drug candidates. and Salts of Weak Electrolytes. The intrinsic solubil-
In this section, the fundamental principles of ioni- ity of a nonelectrolyte is governed by the relative
zation potential and the solubility of weak electrolytes strengths of the intermolecular forces in the solute
and their respective salts are described. Subsequently, (i.e., its solid-state properties) and the affinity of the
the formulation considerations when using buffered solute for the solvent (i.e., the strength of solute-solvent
systems and salts for parenteral delivery and oral interactions) (see Section I.B). This is also the case for
delivery are addressed before finally providing an a weak electrolyte but is complicated by differences in
overview of common salt-screening strategies.
ionization of the solute with changes in pH that, in turn,
A. Solution Behavior of Weak Electrolytes and alter polarity and the affinity of the solute for the
Their Salts solvent. Thus, a weak electrolyte will initially dissolve
to an extent governed by the intrinsic solubility of the
1. Ionic Equilibria. On addition to water, strong
electrolytes such as hydrochloric acid and sodium TABLE 7
hydroxide dissociate almost completely. In contrast, Relative acidity/basicity of molecules with varying pKa values, and
weak electrolytes, such as the weakly acidic and approximate pKa values for common functional groups
weakly basic molecules that constitute most ionizable Acidic Groups pKa Value Basic Groups pKa Value
drugs, dissociate only partially at most pH values. For Sulfonic acids ,1 Aliphatic amines 8–11
a weak acid or weak base, the total concentration in Phosphonates 2 Saturated nitrogen 9–11
heterocycles
solution at a given pH is therefore the sum of the Carboxylic acids 2.5–5 Pyridines 4–6
concentrations of the ionized and un-ionized species. Imides 8–11 Anilines 3–5
Thiols 7–10
The degree of dissociation into the ionized and un- Sulphonamides 8–10 Weak base 4.5–9.5
ionized form is described by the dissociation constant Phenols 8–10 Very weak base 0.0–4.5
(Ka) for the equilibrium reactions shown in eq. 12 for Weak acid 4.5–9.5
Very weak acids 9.5–14
a monoprotic weak acid (HA) (left equation) and weak
336 Williams et al.

un-ionized form. Depending on the pH, a proportion of resulting in precipitation of the solid salt of the acid-base
the dissolved species subsequently ionizes in accordance pairs. The solubility of the salt is therefore the limiting
with the Henderson-Hasselbalch equation (eq. 14), condition to the aqueous solubility of drug in the ionized
leading to further dissolution of solid drug to maintain form when a counterion is present. Where multiple
a constant concentration (i.e., the saturated solubility) counterions are present, the limiting salt solubility
of the un-ionized form. This will continue within an typically reflects the salt with the lowest solubility, or
unbuffered solution until the saturated solubility of the salt formed with the counterion in highest concen-
both un-ionized and ionized species is reached. As the tration if common ion effects are significant. Salt
ionized species is able to form favorable ion-dipole solubility, including aspects of common ion effects, is
interactions with polar solvents, it is more soluble than described in more detail in Section IV.4. Second, classic
the un-ionized species, and therefore, drugs are more ionization texts usually ignore the possibility of ion-pair
soluble in polar solvents at pH values that favor the formation or self-association. Where ion-pair formation
ionized form. The total solubility (S) of a weak acid or self-association leads to the formation of supramolec-
(HA) or weak base (B) at a given pH is therefore the ular complexes in solution (including multimers,
sum of the solubility of the ionized and un-ionized micelles, and liquid crystalline species), apparent solu-
forms (Yalkowsky, 1999). bility values may be much greater than anticipated
  (Serajuddin et al., 1987; King et al., 1989; Kriwet and
S5½A 2  1 ½HA S5½B 1 BH 1 ð15Þ Muller-Goyman, 1993; Fini et al., 1995).
Exemplar pH solubility profiles for a weak acid and
Combinations of eq. 14 and eq. 15 generate the a weak base, and the salts formed by association with
following solubility equations for weak acids (left, eq. a corresponding basic or acidic counterion, are pre-
16) and weak bases (right, eq. 16) where total drug sented in Fig. 7. The total drug solubility (S in eq. 16)
solubility at a particular pH is a function of the of a weak base is low at high pH where the un-ionized
solubility of the un-ionized form (S0) and the pKa of the form (B) predominates (region 1 of the pH solubility
weak acid or base. profile for a weak base). Where the solubility of the
h i h i ionized drug species (BH1) is significantly in excess of
S5S0 1 1 10ðpH 2 pKa Þ S5S0 1 1 10ðpKa 2 pHÞ ð16Þ that of the un-ionized form, B (which is typical),
decreasing pH results in an increase in the proportion
From eq. 16, it is evident that for a monoprotic weak of BH1 and an increase in total solubility. The increase
acid or base, changes in pH of 1 unit away from the pKa in solubility becomes particularly significant from pH
lead to 10-fold (1 log unit) differences in solubility. For values of approximately 1 pH unit above the pKa
diprotic species, similar relationships may be devel- (i.e., where the degree of dissociation increases beyond
oped (eq. 17) where total solubility is dependent on the ;10%) to pH values below the pKa where, as described
pKa of two ionization centers (Yalkowsky, 1999). Ac- previously, solubility increases 10-fold for every unit
cordingly, diprotic species are expected to show a 100- pH decrease (for a weak base). At a critical point (and
fold (2-log unit) change in solubility with a change in critical pH; denoted pHmax), the equilibrium solubility of
pH of one unit (Garren and Pyter, 1990; Venkatesh the ionized species (as the salt) is reached, and no
et al., 1996). The top equation is for a weak diprotic further increases in solubility are seen with decreasing
acid, and the bottom equation is for a weak diprotic pH. This limiting solubility is set by the solubility
base: product of the salt (Ksp):

    Ksp  
S5S0 1 1 10ðpH 2 pKa;1 Þ 1 10ð2pH 2 pKa;1 2 pKa;2 Þ BH 1 X 2 ⇌ BH 1 1 ½X 2  ð18Þ
  ð17Þ solid
S5S0 1 1 10ðpKa;1 2 pHÞ 1 10ðpKa;1 2 pKa;2 2 2pHÞ
When the concentration of solid phase (i.e., salt) is in
excess and, therefore, approximately constant:
The solubility equations for weak electrolytes shown
 
in the preceding should be viewed with two significant K sp 5 BH 1 ½X 2  ð19Þ
caveats. First, eq. 16 and eq. 17 suggest that drug
solubility changes indefinitely with changing pH. In region 2 of Fig. 7 (i.e., at the solubility limit), ex-
Ultimately, however, the achievable solubility is cess solid phase is present as the salt, and the satu-
regulated in most cases by the presence of counterions rated species in solution is largely the ionized form.
that may be intentionally present (in the case of a salt) At pH values ,pHmax, the total solubility of a weak
or unintentionally present (in the case of ions present base is given by [BH1] plus a small (and decreasing)
in blood, GI fluids, buffers, and so forth). In this case, quantity of B (free base) (Kramer and Flynn, 1972;
the total solubility is regulated by the limit of solubility Serajuddin, 2007). The total solubility in region 2 of
of the ionized weak acid or base in the presence of the Fig. 7 is given by eq. 20. for weak acids (top equation)
corresponding basic or acidic counterion, ultimately and weak bases (bottom equation).
Strategies for Low Drug Solubility 337

Scheme 1. Phase solubility scheme for salt formation between a weak


base (B) with a strong acid (X-). Adapted from Pudipeddi et al. (2002).

In a solution of a monovalent salt, dissociation leads


to equal concentrations of [BH1] and [X2]. From eq. 19,
the maximum solubility of [BH1] (i.e., the limiting
value for salt solubility) is equal to the square root of
the solubility product (√Ksp). Ksp is a constant for
a particular salt and defines the maximum concentra-
tion of drug ions and corresponding counterions that
a solvent can maintain in solution (Bhattachar et al.,
2006). Essentially, the higher the Ksp value, the greater
the solubility of the salt. Ksp is therefore a critical
parameter in understanding the solubility of salts or the
Fig. 7. The pH solubility profiles for (A) a weakly basic and (B) a weakly solubility of ionized drugs where salt formation in situ is
acidic compound. In region 1, the un-ionized free acid or free base is likely. Unfortunately, techniques that accurately predict
present as the solid phase and as the saturation or equilibrium species in
solution. Total solubility is a function of both un-ionized and ionized Ksp do not exist, largely as a result of the unpredict-
species (eq. 16), and at pH values greater than (for a weak acid) and less ability of the nature of the crystal lattice. Screening
than (for a weak base) pKa, solubility increases because of increased
quantities of the ionized species in solution. Solubility is ultimately
programs that are able to rationally and rapidly assess
limited in the “salt plateau” region (region 2) by Ksp (eq. 19). In region 2, the limit of solubility of different salt forms are therefore
the salt is present as the solid phase, and the ionized form is the saturated critical to support effective development programs, and
or equilibrium species in solution. At pH values greater than (for a weak
acid) and lower than (for a weak base) pHmax, the total solubility decreases are described in more detail in Section IV.E.2.
slightly because of a reduction in the quantity of free acid or free base in 3. Factors Affecting Salt Formation at pHmax.
solution (eq. 20). Where pH change continues as a result of the further
addition of acidic or basic counterion, salt solubility in region 2 may The pHmax defines the limiting pH above which (for a
decrease much further (dotted line) as a result of the common ion effect. weak acid) and below which (for a weak base) the solid
 phase constitutes the salt rather than the free acid or
½H3 O 1  h i
2
S5½A  1 1 5½A 2  1 1 10ðpKa 2 pHÞ free base. To facilitate salt formation, the bases or
Ka acids that are used to provide the counterion must be
 ð20Þ
1 ½K a   1
h ðpH 2 pKa Þ
i sufficiently strong to shift solution pH past pHmax (into
S5½BH  1 1 5 BH 1 1 10 region 2 of Fig. 7). In general, salts of weak bases will
H3 O 1 form more readily with more strongly acidic counter-
ions such as hydrochloride compared with more weakly
Scheme 1 summarizes the solid-solution equilibria acidic counterions such as citrate. As a general rule,
for a weak base (B), where the counterion is provided salt formation will occur only between a weak acid or
by a strong acid (HX) to form a salt (BH1X-)solid. This weak base and counterion if the difference in pKa for
scheme brings together the relationships between the two is greater than 2 (Pudipeddi et al., 2002; Stahl
dissociation (described by Ka and the dominant factor and Wermuth, 2002; Black et al., 2007). It is also
in region 1 of Fig. 7) and the equilibrium solubility of evident that conditions that cause a shift in pHmax will
the salt (described by Ksp and the dominant factor in change the ease of salt formulation. Thus, conditions
region 2 of Fig. 7). In some cases, additional equilibria, that result in an increase in pHmax for a weak base
including the potential for hydrate formation or self- and a decrease in pHmax for a weak acid will assist in
association, may be present but have been omitted salt formation by reducing the pH shift that must be
here for the sake of clarity. generated by the counterion to move into region 2 of
338 Williams et al.

Fig. 7. For example, decreasing the pHmax for a weak be taken at pH values that ensure that the solid phase
acid will allow salt formation with a more weakly basic in equilibrium with solution is the salt and not the free
counterion, thereby increasing the possible choices of acid or base. As such, solubility measurement should
counterion. be taken at pH values above (for weak acids) or below
At the pHmax value, the salt and the free acid or free (weak bases) the corresponding pHmax. If this condition
base are present in the solid state, and this situation is is not met, the solution is not saturated with respect to
unique to this transition point in the pH-solubility drug in the ionized form, leading to an underestima-
profile. Under these circumstances, eq. 16 and eq. 20 tion of the saturated salt solubility. It is advisable
are both applicable. Setting eq. 16 equal to eq. 20 and during solubility measurements to determine both the
solving for pH (i.e., pHmax) results in eq. 21 and a pH of the solution and to characterize the nature of
relationship that provides an indication of the de- the solid phase (e.g., using XRPD). The presence of
pendence of pHmax on pKa, the intrinsic solubility of the salt alone in the analyzed solid provides evidence of
un-ionized free acid or free base (S0) and Ksp (Bogardus location within region 2 of the pH-solubility phase
and Blackwood, 1979; Serajuddin, 2007). Note that the diagram (Fig. 7) and, therefore, that the data obtained
equation on the left applies for weak acids while the reflect the true solubility of the salt. Conversely, the
equation on the right applies for weak bases: presence of free acid or free base suggests that insuf-
pffiffiffiffiffiffiffiffi ficient salt has been added to reach the pHmax and that
K sp S0 the system has equilibrated in region 1 (Alsenz and
pHmax 5pKa 1 log pHmax 5pKa 1 log pffiffiffiffiffiffiffiffi
S0 K sp Kansy, 2007; Hasa et al., 2011). Under these circum-
ð21Þ stances, the addition of larger quantities of salt is
required to provide further counterion, which in turn
From eq. 21, it is clear that the value for pHmax is results in a pH change (Pudipeddi et al., 2002).
dependent on the properties of the drug namely, Typically, a relatively high excess of salt is required
intrinsic solubility (S0), relative acidity/basicity (pKa) to ensure that the pH approaches pHmax. Determina-
and the solubility of the salt (Ksp). The effect of changing tion of the solubility of a salt in a buffered solution will
these three terms on pHmax is illustrated for a typical reveal only the solubility of the un-ionized plus ionized
weak base in Fig. 8. Increasing the pKa (i.e., increasing form of the drug at that particular pH, and where the
basicity) (Fig. 8A), increasing S0 (i.e., increasing in- pH is above pHmax (for a weak base) or below pHmax
trinsic solubility) (Fig. 8B), and decreasing Ksp (i.e., (for a weak acid), the solubility of the drug is
forming salts with lower solubilities), all assist salt dependent only on pH and the solubility of the ionized
formation by increasing pHmax (realizing that lower form (eq. 16) and is independent of the counterion. This
solubility salts may not be preferred in many cases). provides no indication of Ksp, and therefore, no indi-
4. Determinants of Salt Solubility. The maximum cation of the solubility limit for the salt. The reader is
solubility of a salt is the boundary condition for in vivo directed to a recent article by Murdande et al. (2011a)
solubility and dissolution and, therefore, the parame- for a more detailed description of the factors to be
ter against which alternate salts are screened (at least considered when measuring the equilibrium solubility
for their impact on solubility). The accurate determi- of crystalline solids (and applicable here to salts).
nation of salt solubility, however, is not a trivial task b. Choice of Counterion to Maximize Salt Solubility.
due to the potential for interconversion between the Table 8 provides a list of common counterions that
salt form and the free base or free acid (see Section IV. have been used to isolate salts of weakly acidic and
D.1.c), and the impact of a number of system variables. weakly basic drugs for the development of oral and
These are described below. parenteral drug products. An indication of the fre-
a. pHmax. To ensure that solubility data reflect the quency of use is also provided (Paulekuhn et al., 2007).
equilibrium solubility of the salt, measurements must As Ksp is counterion specific, the solubilities of salts

Fig. 8. The effect of (A) pKa, (B) intrinsic drug solubility (S0) and (c) salt solubility (Ksp) on the position of the pHmax of a weak base. Salt formation of
a weak base becomes more attainable (an increase in pHmax) with increasing pKa of the base, an increase in intrinsic solubility, or a decrease in salt
solubility. Adapted from Pudipeddi et al. (2002).
Strategies for Low Drug Solubility 339
TABLE 8
List of acidic and basic counterions commonly used in pharmaceutical formulations
Frequency (%)b Frequency (%) b

Acidic pKaa Basic pKaa


Oral Parenteral Oral Parenteral

Iodide (most acidic) #26 0.3 1.0 Potassium (most basic) ;14 13.3 1.0
Bromide #6 4.1 4.3 Sodium ;14 65.3 85.4
Chloride 26 56.6 53.4 Calcium 12.6 12.0 2.9
Sulfate 23 7.5 8.2 Magnesium 11.4 2.7
Ethandisulfonate 22.1 0.3 0.5 Cholinate . 11 1.3
Tosylate 21.34 0.3 0.5 Lysine 10.93 1.0
Nitrate 21.32 0.6 0.5 Diethylamine 10.93 1.0
Mesylate 21.2 4.4 3.9 Meglumine 9.5 4.9
Napsylate 0.17 0.6 Diethanolium 9.28 1.0
Trifluoroacetate 0.52 0.5 Procaine 9.0 1.0
Besylate 0.7 0.6 1.4 Tromethamine 8.02 2.7 1.0
Oxalate 1.271 0.3 Piperazine 5.68 1.3
Laurylsulfate 1.3 0.3 Benzathine 4.46 1.3 1.0
Isethionate 1.66 1.0
Maleate 1.92 6.9 0.5
Phosphate 1.96 2.5 3.4
Camphorsulfate 2.14 0.5
Tartrate 3.02 2.8 3.9
Fumarate 3.03 1.6 0.5
Citrate 3.128 3.4 2.4
Glucuronate 3.18 0.5
Lactobionate 3.2 0.5
Malate 3.459 0.3 0.5
Hippurate 3.55 0.3 0.5
Polygalacturonate 3.6 0.3
Gluconate 3.76 0.3 0.5
Benzoate 4.19 0.3
Succinate 4.207 1.9 0.5
Acetate 4.756 0.9 5.8
Lactate 4.9 0.3 2.9
Pamoate 4.9 0.9 0.5
Gluceptate 0.5
Oleate .4 0.5
Octadecanoate 4.9 0.3
Chlortheophyllinate 5.28 0.3 0.5
a
Value corresponding to the strongest acid/base moiety only. Values taken from Stahl and Wermuth (2002).
b
Values taken from Paulekuhn et al. (2007).

formed with a variety of different counterions can vary interactions in the solid state) and the degree of
significantly. For example, solubility differences of up solvation (DGsolvation). Solvation is typically enhanced
to 100-fold have been described for a series of amine by salt formation since dissociation into ions facilities
salts of diclofenac (O’Connor and Corrigan, 2001), and the formation of ion-dipole interactions with water
differences in solubility of up to four orders of molecules that are more energetically favorable than,
magnitude have been described for salts of ibuprofen for example, hydrogen bond interactions between
(David et al., 2012). Similarly, for the weak base water and un-ionized drug. In contrast, the crystal
enalapril, salts formed with a range of aliphatic or lattice energy is typically increased by salt formation
aromatic carboxylic acids exhibit solubilities ranging via the generation of stronger intramolecular ionic
from 22 mg/ml up to 350 mg/ml (Kumar et al., 2008a). interactions, and this is usually evident in an increase
A better understanding of the factors that affect salt in melting point (Corrigan, 2007; David et al., 2012).
solubility (and therefore counterion selection) may be From eq. 22, the increase in (DGlattice) is expected to be
gained by examination of the determinants of the detrimental to solubility. However, the higher solva-
molar free energy of solution (DGsoln) (Chowhan, 1978): tion energies resulting from ionization typically offset
the energy cost of breaking ionic interactions in the
crystal lattice, leading to a net increase in solubility for
DGsoln 5DGsolvation 2 DGlattice ð22Þ
salts compared with the free acid or base.
DGsolvation 5DGcation 2 DGanion ð23Þ The importance of solvation of both anion and cation
to the overall solvation energy dictates that hydrophilic
Solubility is a function of the free energy gain counterions, by virtue of higher solvation energies,
achieved by solvation of the anion and cation (eq. 23) might be expected to form more soluble salts than
offset by the free energy of the crystal lattice (eq. 22). hydrophobic counterions (Menegon et al., 2011; David
Salt formation has the capacity to alter both the free et al., 2012), and indeed this is often the case. However,
energy of the crystal lattice (DGlattice) (by changing the the choice of counterion affects both solvation and
nature of packing and the strength of intermolecular lattice energy, and as such, definitive structure-
340 Williams et al.

solubility relationships for differing counterions do not maleate, besylate, and tosylate salts, despite each of the
exist (Stahl, 2003; Corrigan, 2007; Serajuddin, 2007). counterions showing very high aqueous solubilities.
Thus, monovalent and divalent alkali metals (K1, Na1, Finally, the solubilities of 22 salts formed between vari-
Mg21, Ca21) are widely used counterions for salt ous p-substituted benzoic acid derivatives and benzyl-
formation for weakly acidic drugs (Table 8), and amine have been shown to vary significantly, despite
increasing valency and decreasing counterion ionic broadly similar energies of solvation (Parshad et al.,
radius might be expected to increase hydrophilicity 2002). Across different salt series, lattice energy effects
and facilitate solvation. However, salts of monovalent on solubility therefore appear to dominate over solvation
and larger cations (of equivalent valency) are typically effects in many cases.
the most soluble (Chowhan, 1978; Anderson and Consistent with the potential importance of differ-
Conradi, 1985; Forbes et al., 1995). For example, Fig. 9 ences in lattice energy on salt solubility, several authors
depicts the solubility of metal salts of 7-methylsulfinyl- have described correlations between the melting point
2-xanthonecarboxylic as a function of pH (Chowhan, of pharmaceutical salts (as a measure of lattice energy)
1978) and illustrates that maximum salt solubility is and saturated aqueous solubility (Anderson and Con-
evident for larger monovalent counterions [i.e., K1 radi, 1985; Gould, 1986; Thomas and Rubino, 1996;
(most soluble) . Na1, . Ca21 . Mg21]. These trends Ledwidge and Corrigan, 1998; O’Connor and Corrigan,
suggest that counterion effects on lattice energy (where 2001; Guerrieri et al., 2010). Figure 10, for example,
larger, less highly charged species reduce intermolec- shows the relationship between melting point and
ular attractive forces and typically result in lower solubility for a range of flurbiprofen salts formed with
crystal lattice energies) are more critical determinants different organic and inorganic amine cations and
of the solubility of a salt than corresponding effects on provides good evidence of a relationship between
solvation (where higher-charge densities are expected decreasing melting point and increasing solubility
to increase solvation). Hydrophilic counterions may (Anderson and Conradi, 1985). The properties of the
therefore show higher solvation energies but result in counterion that affect packing within the salt lattice,
lower solubility (compared with a less hydrophilic and therefore the lattice energy, include symmetry and
counterion) by virtue of higher lattice energies. For size, the ability to form hydrogen-bonds with the drug
example, the lactate and mesylate salts of the (in addition to ionic interactions), and the capacity to
experimental antimalarial drug [a-(2-piperidyl)-3,6- delocalize charge (Chowhan, 1978; Gould, 1986; Par-
bis(trifluoromethyl)-9-phenanthrenemethanol] are shad et al., 2004). For example, salts of the weakly basic
more soluble than salts formed with the freely soluble drug candidate (UK-47880) formed with planar and
hydrochloride counterion (Agharkar et al., 1976), and aromatic acids were found to be highly crystalline and to
saccharin salts of lamotrigine are more soluble than have higher melting temperatures compared with an
those formed with the more hydrophilic succinic acid and oily, low melting salt formed with two aliphatic and
fumaric acid counterions (Galcera and Molins, 2009). more flexible carboxylic acids (citric acid and dodecyl
Kumar et al. (2008a) have also shown that malate salts of benzene sulfonic acid) (Gould, 1986). Malate and
enalapril are more than 10 times more soluble than maleate salts of ephedrine also show higher melting

Fig. 9. The pH-solubility profiles for 7-methylsulfinyl-2-xanthonecarbox-


ylic acid and its salts at 25°C. Symbols represent experimental data. The Fig. 10. The relationship between log Ksp and the melting temperature of
point at which horizontal lines cross the upward solubility curve 1:1 amine salts of flurbiprofen. Basic counterions were 1) 5 2-amino-2-
represent the pHmax values for respective salts. Solid symbols indicate methyl-1,3-propanediol, 2) 5 ammonium, 3) 5 tromethamine, 4) 2-amino-
that the starting material was the acid, and open symbols indicate that 2-methylpropanol, 5) 5 tert-butylamine, and 6) 5 adamantanamine.
the starting material was the salt. Adapted from Chowhan (1978). Adapted from Anderson and Conradi (1985).
Strategies for Low Drug Solubility 341

temperatures compared with fumarate and tartrate pharmaceutical factors, including solubility, but also
salts (Black et al., 2007). This has been suggested to on other properties, such as flow, compression, and
reflect the small size of the malate counterion and the stability. Currently, the ability to predict differences in
symmetrical nature of the maleate counterion, both the solid-state properties of salts isolated with differing
factors that favor hydrogen bonding within the salt counterions is limited by a lack of understanding of the
lattice such that the lattice energy is increased (Gould, factors that control packing arrangements during crys-
1986). Similarly, diclofenac salts formed with the freely tallization and the potential for ionized drug molecules
soluble tris(hydroxymethyl)aminomethane (TRIS) coun- to self-associate in solution. This in turn limits accurate
terion were found to be less soluble than expected, de novo prediction of counterion effects on solubility.
a situation later ascribed to the ability of TRIS to Consequently, it is more typical to select the most
hydrogen bond within the salt lattice (Fini et al., 1996; appropriate salt via a salt-screening process. Salt screens
O’Connor and Corrigan, 2001). In this example, the are described in more detail in Section IV.E.2.
TRIS salt had the highest melting temperature and c. The Effect of Common Ions on Salt Solubility.
lowest solubility of a range of diclofenac salts formed The solubility of a salt is highly sensitive to the
with organic amine counterions and therefore provides presence of common ions. This effect can be rational-
another example of the potential for more polar ized according to the theoretical descriptions of salt
counterions to result in lower salt solubilities via solubility as described in Section IV.A.3, in particular
stronger packing within the crystal lattice (Corrigan, eq. 19 and Scheme 1, where it is apparent that excess
2007). In contrast, increasing the alkyl chain length of counterion will result in suppression of salt solubility
amine counterions led to a decrease in the solubility of to maintain Ksp (Berge et al., 1977; Streng et al., 1984;
salts of several weak acids (ibuprofen, gemfibrozil, Serajuddin, 2007). The presence of common ions has
etodolac), although increasing the alkyl chain length frequently been shown to reduce salt solubility in vitro
on the counterion lowered the salt melting temperature (Miyazaki et al., 1980; Streng et al., 1984; Fini et al.,
(David et al., 2012). The overall effect of counterion 1995; Wang et al., 2002; Bergstrom et al., 2004; Li
polarity on solubility therefore reflects the net effect of et al., 2005a; Larsen et al., 2007), and in some cases
changes to both lattice energy (where more polar counter- common ions can suppress the solubility of a salt below
ions typically increase lattice energy, decreasing solubil- that of the free base/acid (Li et al., 2005a). The common
ity) and hydration (where increasing counterion polarity ion effect also has potential implications for parenteral
typically increases hydration and increases solubility). formulations where ions present in various additives
Interestingly, mesylate (methanesulfonate) salts ap- (e.g., buffers, salts) are common to those in the salt.
pear to be finding increasing favor for the delivery of Common ion effects are more critical for drugs of low
weak bases both orally and parenterally (Paulekuhn solubility where small changes to common ion concen-
et al., 2007; Elder et al., 2010; Branham et al., 2012). trations can lead to relatively large compensatory
The wider use of the mesylate counterion in salt changes to drug concentrations (Serajuddin, 2007).
selection strategies for solubility enhancement may be Effects in vivo are also evident where salts come into
in part attributed to its high-charge density but also to contact with high concentrations of common ions. The
its retention of a degree of hydrophobicity, which ap- most familiar example is that of changes to the
pears to affect molecular packing in the crystal lattice. solubility and dissolution rate of hydrochloride salts
As a result, salts formed with the mesylate counterion in the regions of the GI tract that contain endogenous
are likely to show high solvation energies compared chloride ions, such as the stomach (from acid secretion)
with salts formed with other organic counterions and the intestine (due to the presence of sodium
(because of the high-charge density) but lower lattice chloride) (Miyazaki et al., 1981; McConnell et al.,
energies compared with salts formed with inorganic 2008), but this effect is also evident for sodium salts of
counterions (which also show a high charge density). weak acids where concentrations of endogenous so-
For example, mesylate salts of a-(2-piperidyl)-3,6-bis dium ion are high. The use of counterions other than
(trifluoromethyl-)9-phenanthrenemethanol(Agharkar hydrochloride or sodium (i.e., counterions not found in
et al., 1976), RPR200765 (Bastin et al., 2000), and high concentration in vivo) might be expected to reduce
haloperidol (Li et al., 2005b) all showed lower melting the possibility of common ion effects. However, in-
temperatures compared with equivalent hydrochloride terconversion of salt forms during dissolution is
salt forms. The mesylate ion also has a very low pKa possible. For example, Li et al. (2005a) have described
(the lowest of all common organic acidic counterions) a situation where mesylate and phosphate salts of
(Table 8), in turn facilitating salt formation with weak haloperidol that were exposed to high concentrations of
and very weak bases (Stahl, 2003). Mesylate salts are free chloride subsequently converted to the hydrochlo-
also less susceptible to the common ion effect in vivo, ride salt, which showed a slower rate of dissolution
and this is described in more detail in Section IV.D.1.d. resulting from the presence of the common ions.
Finally, the choice of counterion requires apprecia- In vivo salt conversion (most commonly to the
tion of the impact of the counterion on a range of hydrochloride or sodium salt) therefore runs the risk
342 Williams et al.

of a decrease in solubility due to a change in salt form, formation have good solvent properties for un-ionized
and may also increase susceptibility to common ion drug, but they have limited effects on acidity/basicity
effects. These possibilities dictate that the Ksp for (Rubino and Thomas, 1990). Despite the frequency of
chloride and sodium salts is typically measured, use of cosolvents during purification and isolation of salt
regardless of the salt form that is ultimately developed, forms, Serajuddin and Pudipeddi (2002) caution that
to provide an indication of the potential ramification of cosolvent use may result in the isolation of salts that
in vivo conversion. would not ordinarily form in water. Such salts therefore
Interestingly, although salts formed with hydrochlo- are at high risk of conversion to the free acid/base on
ride and sodium ions are perhaps the most susceptible contact with aqueous fluids and moisture (Serajuddin
to common ion effects, they are also by far the most and Pudipeddi, 2002).
common counterions for weak bases and acids, respec-
tively (see Table 8). This situation may be attributed to B. pH adjustment Strategies for Addressing Low
the high acidity/basicity of these counterions and, Drug Solubility
therefore, the potential to form salts with even weakly
ionizable drug compounds (Bastin et al., 2000; Pudipeddi Solution formulations can be readily adjusted to
et al., 2002) and the realization that hydrochloride and a target pH such that a drug candidate is ionized to
sodium salts are often the default option during drug a degree that affords solvation of the target drug dose.
crystallization and purification. In the absence of This approach applies equally to solutions of salts or
obvious problems with factors such as solubility, the corresponding free acid or free base since the
hygroscopicity, stability, and crystallinity, these salt solubility is the same at an equivalent pH and counterion
forms are often progressed, and as such, their high concentration (assuming the solubility limit of the salt
prevalence does not necessarily provide any indication has not been reached).
of advantage with respect to solubility. 1. Buffered Systems Used in Parenteral Formulations.
d. Effect of Organic Solvents on Salt Solubility. The most commonly used pharmaceutical buffers
As a culmination of the synthetic steps required to comprise weak acids (HA) and the conjugate base
obtain a new chemical entity, organic solvents or (A2). Since physiological pH covers mainly neutral and
cosolvent-water mixtures are commonly used to facilitate acidic pH, buffers of weak bases (B) and their conjugate
isolation of a crystalline solid. Isolation as a crystalline acids (BH1) are less frequently used. To exemplify, the
solid assists in purification, and the choice of the initial acid-conjugate base pair of acetate buffer is illustrated
salt form often reflects ease of purification rather than in Scheme 2.
intrinsic downstream pharmaceutical or biopharmaceut- In this system, an increase in hydrogen ion concen-
ical properties. Nonetheless, almost all salt forms are tration is buffered by the conjugate base (CH3COO2),
commonly isolated from organic solvents, mixtures of and the addition of hydroxide ion is similarly neutral-
organic solvents or cosolvent/water mixtures (Serajuddin ized by the acid (CH3COOH). Buffer capacity is widely
and Pudipeddi et al., 2002; Kumar et al., 2007). The used to measure the ability of a buffer to resist a change
organic solvents (water miscible and immiscible) that in pH and is a measure of the quantity of acid/base
may be used for developing drug products are restricted required to change the pH of a buffered solution by 1 pH
by safety and toxicology issues; a comprehensive review unit. Increasing buffer concentrations leads to increases
of the use of nonaqueous solvents for oral drug products in buffer capacity, and at a constant buffer concentra-
was recently published (Pole, 2008). tion, maximum buffering capacity is achieved when the
In addition to changing bulk solubility properties solution pH is close to the pKa of the buffer.
(and promoting crystallization), the presence of cosol- Details of a wide selection of buffers have been
vent can significantly affect the process of salt forma- previously published (Flynn, 1980; Merck-Index, 2006).
tion. For example, in less polar environments (e.g., in Buffer systems most commonly applied to parenteral
the presence of a cosolvent), the ionized form of the drug pharmaceutical formulations and their effective buffer
and counterion is less favored. As such, weak acids and pH range are listed in Table 9 (Lee et al., 2003; Li and
bases become weaker, increasing pKa in the case of Zhao, 2007; Nema and Brendel, 2011). The most
weakly acidic drugs and acidic counterions or decreas- effective buffer range is usually limited to one pH unit
ing pKa in the case of weakly basic drugs and basic either side of the pKa, although buffer combinations
counterions. Under these conditions, the use of solvents may be used to extend the buffer range.
or cosolvents disfavors salt formation by reducing the
ability of the acidic or basic counterion to shift the pH to
pHmax (Tarsa et al., 2010). In contrast, where drug and
counterion pKa remain largely unchanged, solvents and
cosolvents promote salt formation via increases in the
solubility of the un-ionized form (Fig. 8C) (Kramer and
Scheme 2. Dissociation of a commonly used (acetate) buffer.
Flynn, 1972). Ideal solvents and cosolvents for salt
Strategies for Low Drug Solubility 343
TABLE 9
Buffers widely used in parenteral formulations, including details of their buffering range and concentration used in some example
marketed formulations

Buffering Agenta pKa(s) Most Effective pH Rangeb Example Formulations; Buffer Concentration (M);
Route of Administration

Maleic acid 1.9, 6.2 2–3 Librium; 0.14 M i.v.


Tartaric acid 2.9, 4.2 2.5–4 Priscoline; 0.04 M i.v.
Lactic acid 3.8 3–4.5 Ciproxin; ;0.012 M i.v. infusion
Citric acid 3.1, 4.8, 6.4 3–7 Amikin; ;0.13 M i.v. or i.m.
Acetic acid 4.75 4–6 Novantrone; ;0.08 M i.v. infusion
Yutopar; ;0.07 M i.v. infusion
Sodium bicarbonate 6.3, 10.3 4–9 Ivanz; 0.03–0.5 M i.v. or i.m.
Sodium phosphate 2.2, 7.2, 12.4 6–8 Coumadin; ;0.04 M i.v.
Zantac; ;0.02 M i.v. or i.m.
Norcuron; ;0.012 M i.v.
a
Selected LD50 values (i.v. administration): Acetic acid: 525 mg/kg (mouse), sodium acetate: 1195 mg/kg (mouse), citric acid: 330 mg/kg (rabbit), sodium (mono) citrate: 379
mg/kg (rabbit), trisodium citrate: 143 mg/g (rat).
b
Summarized from Li and Zhao (2007), Lee et al. (2003) and Nema and Brendel (2011).

During parenteral administration, pain on adminis- dose. For example, attempts to formulate a drug
tration is likely when using solutions buffered to high candidate that was a weaker acid than naproxen (e.g.,
or low pH. Formulations are therefore typically limited a compound with pKa 5 7), but of equivalent intrinsic
to a working pH range of 4–9, although a wider pH solubility, would provide a maximum concentration of
range (2–9) may be used acutely in a preclinical setting only ;1.6 mg/ml in bicarbonate buffer (pH 9). In this
(Li and Zhao, 2007). Extremes of pH may also lead to case, pH adjustment alone is unlikely to offer a viable
chemical instability, and in these cases, drug stability solubilization strategy, and a combination of strategies
should be monitored in the formulation for the would need to be explored, such as the addition of
intended period of use (Pasini and Indelicato, 1992; a cosolvent, surfactant, or cyclodextrin. The advantages
Islam and Narurkar, 1993; Carstensen, 1995; Zhu (and complications) associated with these combination
et al., 2002; Chen et al., 2006). Buffer concentrations approaches are discussed in the following section, and the
normally exceed 0.05 M and are generally less than 0.5 benefits of cosolvents, surfactants; cyclodextrins as in-
M (Wang and Kowal, 1980; Sinko, 2006). Increasing dividual solubility enhancers are described in more detail
buffer capacity does not increase drug solubility but in Sections VI, VII, and VIII, respectively.
can prove more effective in maintaining drug in 2. Impact of Cosolubilizers and Electrolytes on pH-
solution on dilution or with the addition of a formula- Mediated Solubilization.
tion additive. High buffer concentrations, however, a. pH Adjustment and Cosolvents. Cosolvents—including
increase the risk of pain on injection and in high propylene glycol, ethanol, PEG, and DMSO—increase
volume may alter physiological levels of biologically the solubility of nonpolar molecules in polar solvents
important buffer components such as lactate (Nema (such as water) by reducing solvent polarity (see
and Brendel, 2011). The buffer will also contribute to Sections VI and VI.D.1). The presence of a cosolvent
the overall ionic strength of the formulation, and high may therefore provide additional solubilization for
buffer concentrations can promote salting out of drug solution formulations where pH manipulation is in-
from solution (McDevit and Long, 1952; Raghavan sufficient. However, cosolvents also alter the strength of
et al., 1996; Yalkowsky, 1999). the buffers that are included in a pH-cosolvent
A pH adjustment is a simple and powerful mecha- combination formulation, potentially altering their
nism for solubility adjustment and, as such, is typically application (Narazaki et al., 2007b). For example, acidic
the first approach for solution formulations of ionizable buffers are less acidic in the presence of cosolvents since
drugs. For example, naproxen is a weak acid with the less polar conditions favor the un-ionized form of the
a pKa of 4.57 and intrinsic solubility (of the free acid) of buffering species, resulting in an increase in the
16 mg/ml (Fini et al., 1995). As a comparatively “strong” apparent pKa (Rubino, 1987; Yalkowsky, 1999). Simi-
weak acid, the low solubility of naproxen can be in- larly, decreasing concentrations of the ionized species in
creased dramatically in a pH range acceptable for basic buffers leads to a decrease in pKa and a reduction
parenteral administration. Indeed, solution concentra- in basicity. As acids donate protons, they increase the
tions of ;45 mg/ml might be expected in pH 8 number of free ions in solution; therefore, the pKa of
phosphate buffer and ;450 mg/ml in pH 9 sodium acidic species is more sensitive to changes in solvent
bicarbonate buffer. Under these circumstances, pH polarity compared with bases, which have no effect on
adjustment could be used to deliver a wide range of the concentration of ions in the system (Rubino, 1987;
doses (Lee et al., 2003). In some cases, and in particular Yalkowsky, 1999; Narazaki et al., 2007b). The effect of
for weak(er) acids of higher pKa and weak(er) bases with cosolvents on the pKa of both acids and bases is
lower pKa values, pH adjustment alone may be in- therefore predictably more enhanced with decreasing
sufficient to allow the administration of an appropriate polarity of the cosolvent (i.e., reduced dielectric
344 Williams et al.

capacity). The potential outcome of this effect is that the recent study of the solubility properties of the weak
addition of a cosolvent to a buffer system may alter its acid phenytoin in a formulation containing 50 mM
capacity to maintain pH by causing a shift in pKa of the carbonate buffer (pH ;9.8) with and without 15%
buffering species. The effect is further complicated by ethanol. After dilution with phosphate buffer (pH 7.4),
the ability of cosolvents to shift the pKa of weakly acidic formulations containing ethanol showed a greater shift
and weakly basic drugs such that there is a reduced (decrease) in pH as a result of the increased pKa of the
fraction of ionized drug at a given pH (Kramer and buffer in the presence of the cosolvent, in turn leading
Flynn, 1972; Narazaki et al., 2007b). to increased drug precipitation in comparison with
Importantly, however, the net effect of using a cosol- ethanol-free formulations (Narazaki et al., 2007b).
vent with pH manipulation is usually an increase in b. pH Adjustment and Strong Electrolytes. Strong
total drug solubility. This reflects the fact that changes electrolytes, such as sodium chloride, are often added to
in dissociation are usually offset by cosolvent-mediated parenteral formulations to ensure isotonicity. In contrast
enhancements in the solubility of the un-ionized form. to the effect of cosolvents, the addition of polar species
Cosolvents may in some cases also increase the such as sodium chloride increases the polarity of the
solubility of the ionized species. Table 10 describes solvent, causing a reduction in the solubility of the un-
the effect of the addition of a range of cosolvents on the ionized form and an increase in the solubility of the
solubility of a weak base across a wide pH range (1–7). ionized form (effectively increasing acid/base strength)
In all cases, cosolvent addition increases solubility, and (Windheuser and Higuchi, 1963; Flynn, 1980; Yalkow-
the maximum solubility is evident at the lowest pH sky, 1999). Electrolytes therefore decrease the pKa of
(reflecting the largest ionized fraction) (Jain et al., acidic buffers and increase the pKa of basic buffers, with
2001). However, the relative gain that can be attrib- once again the effect being more pronounced for acids.
uted to the cosolvent progressively decreases with There is also the potential for electrolytes to modify
decreasing pH. This reflects the decreased solubility solubility via the common ion effect or the formation of
advantage provided by cosolvents for increasingly salts of lower solubility.
ionized species. c. pH Adjustment and Surfactants. Surfactants usu-
Since the addition of a cosolvent to a pH-buffered ally enhance drug solubility via solubilization in the
solution leads to a net increase in drug solubility, this hydrophobic micelle core (see Section VIII). Consis-
approach has been widely implemented in marketed tent with the effect of cosolvents on buffered solutions,
parenteral formulations, including chlordiazepoxide the beneficial effects of surfactants are therefore
[Librium; Hoffmann-LaRoche, Inc., Nutley, NJ (pro- progressively lost with increasing drug ionization.
pylene glycol 20%, pH 3)], dihydroergotamine [DHE- This is also evident in Table 10, where decreasing pH
45; Novartis AG, Basel, Switzerland (15% glycerin, (and therefore increasing ionization for a weak base)
6.1% ethanol, pH 3.6)], fenoldopam [Corlopam; Ben leads to a decrease in the advantage provided by
Venue Laboratories, Inc., Bedford, OH (propylene micellar solubilization in a 20% Tween 80 solution
glycol 50%, pH 3)], oxytetracycline [Terramycin; Pfizer (Jain et al., 2001). Nonetheless, the net effect is again
Labs, NY (propylene glycol 65–75%, pH 3)], pentobar- an increase in total drug solubility in the presence of
bital [Nembutal; Lundbeck, Inc., Deerfield, IL (pro- both surfactant and reduced pH since both approaches
pylene glycol 40%, ethanol 10%, pH 9.5)], and are beneficial to solubilization. An additional ramifi-
phenytoin [Dilantin; Park-Davis, NY (propylene glycol cation of the combination of pH and surfactant ap-
40%, ethanol 10%, pH 10–12.3)] (Strickley, 2004). proaches is the realization that a pH change can affect
One concern with the use of pH-cosolvent combina- surfactant properties. Thus, changes to pH can sig-
tions is the potential for an increased risk of drug nificantly alter the properties of ionic surfactants
precipitation on dilution resulting from a reduction in (including solubility, critical micelle concentration,
buffer capacity. This is well exemplified by a relatively and solubilization capacity). However, ionic surfac-
tants are in general more toxic than nonionic surfac-
TABLE 10
tants (Strickley, 2004) and are therefore rarely used
The effect of cosolvent and surfactant addition on the solubility of a weak in parenteral formulations. In contrast, nonionic surfac-
base at differing pH tants are more commonly used and have self-association
pH 7a pH 2a pH 1a properties that are more independent of pH. The
mg/ml potential for pH to affect the solubilization of drugs in
Buffer alone 0.00003 0.024 0.210 micellar formulations is described in more detail in
20% Ethanol 0.003 (100) 0.047 (2.0) 0.848 (4.0) Section VII.D.2.
20% DMSO 0.002 (66.7) 0.021 (0.9) 0.511 (2.4)
20% PEG 400 0.006 (200) 0.028 (1.2) 0.702 (3.3) d. pH Adjustment and Cyclodextrins. Many non-
40% Propylene glycol 0.017 (567) 0.119 (5.0) 1.114 (5.3) polar drugs are readily solubilized by cyclodextrins via
20% Tween 80 1.8 (60 000) 1.817 (75.8) 2.948 (14.0)
the formation of an inclusion complex within the
Data taken from Jain et al. (2001). hydrophobic cyclodextrin core (see Sections VIII and
a
Values in parentheses represent the ratio of solubility in the presence of
cosolvent/surfactant to solubility in buffer alone. VII.C). Given the hydrophobic nature of this interaction,
Strategies for Low Drug Solubility 345

ionized drugs typically have lower cyclodextrin binding


constants (Ki) than do the un-ionized species (Ku) (Cirri
et al., 2006). A pH adjustment to achieve conditions
where the ionized species is more dominant therefore
reduces the efficiency of cyclodextrin complexation, and
the Ku:Ki ratio has been used to describe the loss of
binding affinity resulting from increasing drug ioniza-
tion (Okimoto et al., 1996). However, in a fashion
similar to that described for surfactants or cosolvents,
formulations using pH manipulation in combination
with cyclodextrins typically retain improved solubiliza-
tion capacities compared with pH manipulation or
cyclodextrin addition alone (Li et al., 1998a, 1999c; Ni
et al., 2002; Ran et al., 2005; He et al., 2006). This Fig. 11. Effect of lowering pH (increasing acidity) on the dissolved drug
reflects a greater gain in solubility via pH shift concentration of pH-adjusted formulations containing a weak base
compared with the decrease in efficiency of cyclodextrin relative to its solubility curve. Points on the dilution curves at which
theoretical drug concentrations exceed maximum solubility (labeled X
complexation. The gain in solubility with a change in pH and Y) indicate areas of precipitation risk. S0 is the intrinsic solubility of
is a combination of the higher solubility of the ionized the free un-ionized base.
form of the drug and the potential complexation of
ionized drug. Even though the binding constant for the described. Figure 12 shows the effect of diluting
ionized drug species is usually lower than that of the un- formulations of phenytoin (a weak acid) buffered to
ionized species, the total quantity of complexed ionized pH ;10. The x-axis shows dilution in terms of the
drug is typically greater than that of the un-ionized formulation fraction (ff). The ff is defined by eq. 24,
form because the concentration of ionized drug in where Vf is the volume of the formulation and VDM is
solution is much higher. Interestingly, Kim et al. have the volume of the dilution medium, in this case, pH 7.4
shown that the mesylate salt of ziprasidone dissociates Sorensen’s phosphate buffer (SPB) (Narazaki et al.,
incompletely in solution (providing additional evidence 2007a):
of the possibility of stable ion pairs in solution) and also
Vf
that it is the ion pair that preferentially interacts with ff 5 ð24Þ
the cyclodextrin rather than the ionized drug (Kim Vf 1 VDM
et al., 1998). The nature of the counterion used during
salt formation may therefore play a role in drug- Figure 12A is analogous to Fig. 11 and compares the
cyclodextrin binding (Kim et al., 1998; Mura et al., susceptibility to precipitation on dilution of two
2001b). formulations containing either low (0.5 mg/ml) or high
Where the cyclodextrin itself is ionized, the potential (1.0 mg/ml) drug concentrations. As expected, the onset
for enhanced complexation through ion pairing may of precipitation occurs at a lower dilution factor for the
provide additional benefits. This has been observed more concentrated formulation. Figure 12, B and C
with the weak bases prazosin, papaverine, and mico- shows the importance of buffer type and buffer
nazole with anionic sulfobutylether-b-cyclodextrin concentration on the potential for drug precipitation
where the binding constant for the ionized form of on dilution. In this example, carbonate buffer is more
these drugs (Ki) was similar to that of the un-ionized effective in preventing precipitation compared with
forms (Ku) (Okimoto et al., 1996; Zia et al., 2001). phosphate buffer as a result of the superior buffer
3. Effects of Dilution on Drug Solubilization by pH capacity of carbonate buffer over the investigated pH
Adjustment. The effect of a pH change and dilution on range (Narazaki et al., 2007a). Therefore, there is
the solubilization capacity of an acid-adjusted (i.e., not a need to consider not only the ideal pH range in which
buffered) solution of a weak base is illustrated a buffer can operate but also the buffer capacity within
graphically in Fig. 11. In the absence of buffer, dilution this range. In this case, phosphate buffer lacks
is expected to increase pH and, therefore, decrease the sufficient buffer capacity between the pH of the
maximum solubility of the drug. The drop in solubility dilution medium (pH 7.4) and the initial pH of the
is linear when plotted against [H1] (but exponential formulation (pH 10), and consequently the shift in pH
when plotted against pH). Where dilution curves are in phosphate buffer after dilution is greater than that
above the solubility curve (between points X and Y), in carbonate and leads to drug precipitation. The
the solution is supersaturated and at risk of precipitation risk on dilution of buffered solutions may
precipitation. also be predicted; however, predictive models are likely
For a buffered solution, dilution is expected to reduce to portray the worst-case scenario and do not take into
buffer capacity such that pH may increase and drug consideration the possibility of periods of supersatura-
solubility may decrease in a manner to that already tion before precipitation (Narazaki et al., 2007a).
346 Williams et al.

associated with precipitation may take 24 h to materi-


alize) and largely qualitative unless efforts are made to
quantitate precipitation by monitoring, for example, an
increase in temperature at the site of administration
(Ward and Yalkowsky, 1993; Johnson et al., 2003).
Alternatively, in vitro precipitation models may be used
(Yalkowsky et al., 1983, 1998; Li et al., 1998b; Dai,
2010). Briefly, simple serial dilutions of the formulation
are generated with a variety of solutions, including
water, saline, dextrose 5%, human plasma, tris buffer,
or Sorensen’s phosphate buffer (SPB) (Reed and Yal-
kowsky, 1985; Yalkowsky et al., 1998) and evidence of
drug precipitation detected visually or spectrophotomet-
rically following filtration (or another appropriate sep-
aration means) of the solutions. Similarly, the dropwise
addition method may be used where small aliquots of
formulation are slowly introduced (approximately every
30 s) into a static or continuously stirred dilution
medium (El-Sayed and Repta, 1983; Li et al., 1998b).
The latter method has the advantage of revealing the
minimum volume ratio of formulation-to-dilution me-
dium that promotes precipitation. Dynamic precipita-
tion methods are designed to mimic better the effects in
vivo via the addition of formulation to a circulating
medium (normally pumped through plastic tubing using
a peristaltic pump) (Yalkowsky et al., 1983, 1998;
Johnson et al., 2003). A major advantage of this
approach is that the effect of injection rate on the
likelihood of drug precipitation can be assessed (Li et al.,
1998b), and although dynamic tests are more complex
than static in vitro tests, they have been shown to better
correlate with in vivo bioavailability data (Cox et al.,
1991; Dai, 2010).
The duration of an in vitro precipitation test is
important since drug precipitation may not be imme-
diate and may occur at varying rates and extents,
Fig. 12. Effect of (A) initial phenytoin concentration, (B) buffer depending on the formulation and the nature of the
composition, and (C) buffer concentration on dissolved concentration of dilution medium. Li and Zhao (2007) proposed that 3–5
phenytoin with dilution using Sorensen’s phosphate buffer. Dotted lines
denote drug concentrations. Gray symbols indicate evidence of drug
min is a sufficient to assess the likelihood of drug
precipitation. (A) Formulation contained 50 mM of carbonate buffer precipitation during static serial dilution tests as an
containing 0.5 mg/ml (triangles) or 1.0 mg/ml (diamonds) of phenytoin, (B) administered formulation is likely to have been
Formulations contained either 50 mM of carbonate (diamonds) or
phosphate buffer (triangles) and 1.0 mg/ml phenytoin. (C) Formulations significantly diluted in this time after circulation with
contained 10 mM (triangles) or 100 mM (diamonds) of carbonate buffer the blood. Also, in many cases, the administered drug
and 1.0 mg/ml phenytoin. Formulation fraction (ff) is defined in eq. 24.
Adapted from Narazaki et al. (2007a).
will bind to components of human plasma (e.g., albumin
and lipoproteins); therefore, in vitro precipitation
methods that use simple solutions and buffers may
a. Methods for Assessing Precipitation Potential for overestimate the extent of drug precipitation (Cox
Buffered Parenteral Formulations. The importance of et al., 1991). The use of more complex dilution media
drug precipitation on injection has led to the develop- (i.e., containing plasma proteins) may serve to alleviate
ment of a number of models to assess precipitation these differences, although conscious omission of these
potential. Drug precipitation can be visually assessed materials provides a more conservative estimate of
in vivo using various animal models, including the precipitation potential.
rabbit ear, rat tail vein, dog femoral vein, or monkey 4. Buffer Systems in Nonparenteral Formulations.
jugular vein, with phlebitis at or near the injection Solution formulations intended for—for example, oral,
site providing indirect evidence of drug precipitation ocular, or nasal administration—also commonly use
(Yalkowsky et al., 1998; Johnson et al., 2003). However, buffers. Similar to parenteral formulations, buffers
these in vivo methods are slow (sometimes symptoms may be used in these systems to alter pH such that
Strategies for Low Drug Solubility 347

conditions favor ionization of the drug (and therefore addition of different salts of the same drug also varies
increased solubility), and tighter control of pH might depending on the acid/base strength of the counterion.
also be used to ensure adequate drug chemical stability In situations where complete dissociation of a salt in
or permeability (Chung et al., 1970; Mitra and solution results in a pH value that is outside the
Mikkelson, 1982; Washington et al., 2000). In ocular acceptable range for parenteral administration, a buffer
systems, pH buffering can also be used to reduce can be added to adjust the pH. However, this approach
irritation associated with the use of solutions contain- may lead to drug precipitation if the use of a buffer to
ing acidic molecules (Suhonen et al., 1995; Vaddi et al., lower the pH reduces solubility (by decreasing drug
2007). Buffers may also enhance drug absorption ionization) below the target drug concentration.
across the cornea by increasing the apparent lip- An additional factor that warrants consideration is
ophilicity of the drug via the formation of a drug- the effect of common ions in the buffer since this may
buffer ion pair (Higashiyama et al., 2004). lead to precipitation of the drug salt. For example, the
The basic principles of pH buffering are largely use of buffers containing sodium to promote the
consistent across all routes of administration, although solubility of sodium salts of weak acids may exceed
the intended route of drug absorption will dictate the solubility product of the salt resulting in drug
buffer type and concentration. The acceptable pH precipitation. Under these circumstances, a potassium-
range in ophthalmic solutions is narrow (e.g., between based buffer may be more appropriate.
pH 6 and 8) so that the pH of the product is consistent
with the pH of lacrimal fluid for maximum comfort D. The Use of Salt Forms to Address Low Aqueous
(Carney and Hill, 1976). Buffer concentrations in Solubility in Oral Formulations
ocular systems are likely to be low to avoid irritancy. Pharmaceutical salts are widely used in oral solution
Lacrimal fluid has very limited buffer capacity but still and suspension formulations. Most commonly, salts
can influence the pH of a topically applied solution if are used to enhance solubility, either directly to facilitate
the applied volumes are low (Longwell et al., 1976). the formulation of a solution or indirectly to enhance the
Buffered solutions may also be used after oral in vivo dissolution of a suspension, tablet, or capsule fill
administration; however, the intrinsic buffering prop- material or to aid in the in vitro reconstitution of, for
erties of GI contents in either fasted or fed conditions is example, antibiotic solutions for pediatric treatments.
likely to be higher than the formulation; therefore, Alternatively, low solubility salts have been used to
reversion to GI pH is likely in most cases. taste-mask some antibacterials (Jones et al., 1969;
Although manipulation of ionization (and therefore Menegon et al., 2011).
solubility) may be achieved using buffers or direct pH The isolation of differing salt forms may also be used
modification in preclinical studies, the more common to improve other pharmaceutical properties such as
approach in the clinic is the use of salts to allow the stability, flow, compaction, and so forth. The effect of
administration of solid dosage forms. These approaches salt formulation and counterion choice on nonsolubility-
are discussed below. related phenomena are discussed in detail elsewhere
(Giron and Grant, 2002; Stahl, 2003; Huang and Tong,
C. The Use of Salts to Address Low Aqueous Solubility 2004).
in Parenteral Formulations 1. The Use of Pharmaceutical Salts to Enhance
Since salts are often the preferred pharmaceutical Dissolution. The potential for salts to dissociate in
form of a drug (because of ease of isolation, speed of water to form ionized species leads to increases in
dissolution, solid-state characteristics, improved proc- solubility compared with the un-ionized free acid or
essability, and chemical stability), parenteral formula- free base. The solubility difference between a salt and
tions are often prepared using buffered solutions of the free base or free acid may be large (.100-fold),
salts, even though similar endpoints may be achieved especially where the drug is highly hydrophobic. From
using the free acid or free base. Simple solutions of the the Noyes-Whitney equation (eq. 9), the greater
salt (in the absence of additional buffer components) are solubility of salts compared with the free acid or base
also possible, although in this case the limited buffer also suggests the potential for increases in dissolution
capacity of the salt usually leads to a solution that is rate, and in general, this is the case. For example,
sensitive to pH shifts and, therefore, precipitation. sodium and potassium salts of ampicillin show supe-
In the absence of an additional buffer, solutions of rior dissolution rates compared with the free acid
a salt formed using a weak acid and strong base will be trihydrate (Hitzenberger and Jaschek, 1974), and the
basic and those from a weak base and strong acid will dissolution rates of hydrochloride and mesylate salts of
be acidic. The pH at the point at which a solution is haloperidol are significantly higher than the free base
saturated with respect to the salt is the pHmax value, under pH conditions reflective of the small intestine (Li
and this varies depending on the nature of the et al., 2005b). Different salt forms also show differences
counterion and the solubility of the formed salt. In in dissolution rate. Monovalent salts of p-aminosalicylic
undersaturated solutions, the pH value after the acid, for example, generate faster intrinsic rates of
348 Williams et al.

dissolution compared with the equivalent divalent salts, point of difference when assessing the dissolution of
and the dissolution rates of the mesylate salt of a salt, rather than a nonionizable drug or a free acid or
haloperidol is higher than that of the hydrochloride salt free base, is the potential for dissolution of the salt to
(Forbes et al., 1995; Li et al., 2005a). The differences in alter the pH in the unstirred layer, which in turn
dissolution rate generally reflect differences in the affects the solubility of the drug (Cs) in this layer. Since
solubility of the salts, and in some cases, it has been by definition, this water layer is unstirred, significant
possible to correlate salt solubility with the intrinsic differences in bulk pH and microenvironment pH at
dissolution rate (Nicklasson et al., 1981; Forbes et al., the surface of a dissolving salt can be generated and
1995; O’Connor and Corrigan, 2001). maintained. It is this pH difference that explains the
The potential for salts to enhance solubility and often higher drug solubility in the unstirred layer
dissolution rate has led to widespread application in (when compared with solubility at bulk pH) and benefits
oral formulations for poorly water-soluble drugs (Stahl to dissolution rate. In essence, the salt has the potential
and Wermuth, 2002), and several examples of improved to buffer the pH of the immediate microenvironment into
oral bioavailability and/or rate of oral absorption which it dissolves, hence the use of the term self-buffering
(i.e., speed of onset) have been described (Nelson, 1957; (Serajuddin and Jarowski, 1985b).
Morozowich et al., 1962; Nelson et al., 1962; Lin et al., The self-buffering capacity of salts is well exempli-
1972; Jaschek, 1974; Meanwell et al., 1992; Kwei et al., fied by comparison of the dissolution profiles of two
1995; Patt et al., 1999; Desjardins et al., 2002; Gwak different haloperidol salts with that of the free base in
et al., 2005; Guzmán et al., 2007; Han and Choi, 2007; Fig. 13 (Li et al., 2005b). Haloperidol is a poorly soluble
Hitzenberger and Kesisoglou and Wu, 2008; Chiang weak base (pKa ;8), and the free base shows a very
et al., 2009; Menon et al., 2009). It is apparent, however, slow rate of dissolution at pH values above ;2 (Fig.
that the degree of ionization (and therefore solubility) of 13A). In contrast, both the hydrochloride (Fig. 13B)
an orally administered weak acid or weak base is likely and mesylate (Fig. 13C) salts show evidence of reason-
to be dictated in large part by the pH of the GI fluids able dissolution at pH 5 and pH 7. The dissolution rate
rather than the intrinsic pH attained by the dissolution of the mesylate salt is some 6-fold higher than that of
and dissociation of the salt of a weak acid or base in the hydrochloride, a difference consistent with the
water. For example, where GI pH is above the pHmax (for relative differences in the solubilities (Ksp) of the two
a weak base) or below pHmax (for a weak acid), solubility salts. The rationale for the greater dissolution rate of
(and therefore potentially dissolution rate) is expected to the two salts at higher bulk pH values has been
be dictated by the effect of pH on ionization of the free explored by estimating the likely pH in the unstirred
acid or base and will therefore be largely independent of water layer. This was achieved by measuring the pH of
salt form. a concentrated slurry of haloperidol or the haloperidol
In contrast, the dissolution rates of pharmaceutical salts in unbuffered dissolution media (Serajuddin and
salts are often significantly higher than equilibrium Jarowski, 1985a,b). This analysis revealed that disso-
solubility measurements at GI pH would suggest. Salts lution of the free base increased the pH in the
therefore commonly provide benefits in in vivo disso- unstirred water layer relative to bulk pH, effectively
lution and absorption, even when equilibrium solubil- suppressing solubility and dissolution rate (Fig. 13D).
ities at intestinal pH suggest little advantage over free In contrast, dissolution of the salts resulted in
acid or free base. A classic example is provided by equivalent pH conditions in the bulk solution and the
doxycycline, which, as a hydrochloride salt, shows unstirred water layer up to ;pH 3 for the mesylate salt
significantly higher dissolution rates than the free base and pH 5 for the hydrochloride salt, beyond which the
at pH 4 and pH 7 despite similar solubility for both pH remained consistent (i.e., was effectively buffered)
forms of the drug under these pH conditions (Bogardus in the unstirred water layer, even with increases in
and Blackwood, 1979; Serajuddin, 2007). This effect bulk pH up to 8 (Fig. 13D). Substitution of saturated
can be ascribed to the self-buffering potential of salts drug solubilities at the effective pH in the diffusion
during dissolution as described below. layer, rather than solubility at bulk pH, subsequently
a. Effect of Self-Buffering on Salt Dissolution. allowed good correlation of drug solubility with
The dissolution rate of a dissolving solid is limited by dissolution rate (Serajuddin and Jarowski, 1985a,b;
drug diffusion across an unstirred layer of water that is Li et al., 2005b). The self-buffering capabilities of salts
present on the surface of every dissolving solid. The therefore provide the potential for faster rates of
rate of dissolution (dW/dt) is a function of the solubility dissolution across a much larger pH range than is
of the drug in the unstirred layer (Cs), the diffusion possible with either the free acid or free base and
coefficient of the drug in the unstirred layer (D), the facilitates, for example, enhanced dissolution of an
width of the unstirred layer (h) and the drug acidic drug in the low pH environment in the stomach
concentration in the bulk solvent (C) (eq. 9). (Mooney et al., 1981; Serajuddin and Jarowski, 1985a;
The dissolution rate is therefore expected to reflect McNamara and Amidon, 1986, 1988; Southard et al.,
the saturated drug solubility (Cs). However, the critical 1992). The dissolution rate of haloperidol-free base and
Strategies for Low Drug Solubility 349

Fig. 13. Effect of dissolution media pH on the dissolution rate of (A) haloperidol free base and (B) hydrochloride and (C) mesylate salt forms, and (D)
the pH at the surface of a solid as a function of bulk pH of the dissolution media. Dissolution values are expressed as means (n 5 6). Dashed lines are
used for pH 7.01 as only data during the first 30 min of dissolution was provided. Adapted from Li et al. (2005b).

the hydrochloride and mesylate salt forms at pH ;1 bulk pH and the potential for drug precipitation.
were all low (Fig. 13). This slow rate of dissolution of Where the pH of the dissolution fluid encourages
the hydrochloride salt form in the acidic media can be conversion of dissolved salt to the free acid or the free
explained by the common ion effect, whereas the slow base, the bulk solution is referred to as a “reactive”
rate of dissolution of the free base and mesylate salt medium (Higuchi et al., 1964). Significant pH change is
may reflect an in situ conversion to the slowly also evident as, for example, drug in solution passes
dissolving hydrochloride salt. The effect of in situ salt from the gastric to intestinal environment.
interconversions and common ions on dissolution of The implications of pH change on drug precipitation
salts is described in more detail in Sections IV.D.1.c in the GI tract have been extensively discussed (Perng
and IV.D.1.d, respectively. et al., 2003; Kostewicz et al., 2004; Gu et al., 2005; Dai
b. Effect of pH Changes on Drug Supersaturation and et al., 2007; Guzmán et al., 2007; Carlert et al., 2010).
Precipitation. Although the self-buffering effect of Many other factors beyond pH may also play a role in
salts is likely to encourage faster dissolution rates, the propensity for salts to precipitate in vivo, including
diffusion of the ionized species across the unstirred in situ conversion to less soluble salts (see Section IV.
water layer ultimately leads to entry into the bulk D.1.c), the common ion effect (see Section IV.D.1.d),
fluids of the GI tract (or dissolution medium in vitro) and differences in the solubilization capacity of the GI
where mixing precludes the maintenance of the tract owing to the presence of bile salt-phospholipid
buffered pH in the unstirred layer. Under these mixed micelles (Kaukonen et al., 2004a; Carlert et al.,
conditions, contact with the bulk solution results in 2010; Bevernage et al., 2011). Figure 14 provides
supersaturation relative to the solubility of the drug at a simple summary of the potential implications of pH
350 Williams et al.

change on the dissolution (and potential precipitation) the stomach or weak bases in the intestine (Miller
of pharmaceutical salts. Issues pertaining to pH et al., 2008; DiNunzio et al., 2010a). Salts selected to
gradients in a single GI compartment (e.g., the self- provide more controlled (i.e., slower) rates of dissolu-
buffering effects of salts) and during GI transit (e.g., tion may also limit the quantity of drug in the
when moving from the stomach to the intestine) are supersaturated state so that the driving force for
described. First, for the salt of a weak acid in the precipitation is reduced.
stomach (Fig. 14A), dissolution is driven initially by c. Potential for Salt Conversion to Un-Ionized Drug/
self-buffering (to more basic pH) in the diffusion layer. Hydrates/Other Salt Forms In Situ. As described in
Ultimately, however, conversion of the ionized drug to the preceding section, several scenarios may be
the free acid is favored in the bulk acidic pH in the envisaged as a dissolved salt is transported along the
stomach, leading to solution conditions that are changing pH conditions of the GI tract, with the
supersaturated and, therefore, prone to precipitation potential for salt conversion to the free acid or free
of the free acid. Where precipitation of the free acid is base in situ. The fundamental origins for such
evident, this may manifest either as a finely divided conversions are illustrated in Fig. 15, which shows
powder dispersed in the stomach fluids or as a poorly the position of pHmax in relation to the physiologic pH
soluble layer of free acid on the surface of the of the GI tract for weak acids of differing acidity (top
dissolving salt particles (Higuchi et al., 1965b; Ser- panels) and for weak bases of different basicity (bottom
ajuddin, 2007). In the latter case, the layer of free acid panels). In general, for salts of weak acids, the risk of
can severely restrict further dissolution of the salt by conversion to the free acid, and therefore the risk of
reducing ongoing hydration (Serajuddin and Jarowski, precipitation, is increased where pH , pHmax. Simi-
1985b; Serajuddin et al., 1986). However, this effect is larly, for salts of weak bases, conversion to the free
expected to be limited as the drug empties from the base and the likelihood of precipitation increase where
stomach into the higher pH conditions of the intestine pH . pHmax (Serajuddin, 2007). Accordingly, salts of
that favor ionization (and dissolution) of either the free very weak acids (i.e., higher pKa values) and very weak
acid or the salt. bases (i.e., lower pKa values) show higher risks of
In the case of the dissolution of a weak base in the conversion to the free acid/base since they have higher
acidic gastric environment (Fig. 14B), gastric pH is pHmax values (for acids) or lower pHmax values (for
expected to favor ionization, resulting in rapid initial bases) compared with salts of stronger acids/bases (see
dissolution. On transfer into the intestine, however, Fig. 15). However, the rate of conversion of a salt to the
the increase in pH favors conversion to the free base, un-ionized species in response to a change in pH
resulting in supersaturation. Several eventualities are remains unpredictable (Agoram et al., 2010).
possible. First, where intestinal absorption is rapid, Depending on the availability of counterions in
the drug may be absorbed before precipitation. Alter- solution and the Ksp of the resulting salt complex,
natively, precipitation of the free base may occur either drugs administered as one salt form may also pre-
as the crystalline or amorphous form. If dissolution cipitate out of solution as a different salt, resulting in
was incomplete in the stomach and the precipitated marked changes in properties and behavior (Li et al.,
free base forms a poorly soluble layer on the surface of 2005a; Serajuddin, 2007). Such salt interconversions
dissolving salt particles, dissolution in the intestinal are likely where counterions are present in reasonable
environment will slow significantly. Where precipita- concentration and have the potential to form a salt
tion results in phase separation of either crystalline or with a lower Ksp. Salt interconversions may be further
amorphous drug in the GI fluids, the likelihood of exacerbated by the common ion effect. For example,
redissolution and absorption is dictated by the disso- conversion of salts of weak bases (and in particular
lution rate of the particles formed. Fortunately, pre- very weak bases) to the hydrochloride salt in the
cipitation from supersaturated solutions often leads to stomach may render an orally administered salt form
the generation of particles with small particle sizes or sensitive to a chloride common ion effect (even though
amorphous content, both conditions that favor redis- the originally administered salt was not a hydrochlo-
solution, even from the free base. ride). In situ conversions of this type have been
Additives may also be incorporated into the dosage reported during in vitro dissolution of mesylate and
form in an attempt to stabilize supersaturation. For phosphate salts of haloperidol in the presence of
example, various water-soluble polymers have been chloride ions, although in this case enhancing the
used to stabilize the supersaturated solutions formed dissolution rate of the original salt forms (by increasing
by self-buffered dissolution of salts of weak acids in the available drug surface area) was able to reduce in situ
stomach (Guzmán et al., 2004, 2007; Brouwers et al., conversion (Li et al., 2005a).
2009) and on gastric emptying of solutions of weak Issues surrounding hydrate formation are most
bases (e.g., itraconazole) (van Speybroeck et al., 2011). commonly observed on storage, where conversion of
Alternatively, enteric materials may be used to prevent the anhydrous salt to the hydrate typically results in
the release and precipitation of salts of a weak acid in a reduction in solubility of the drug product. In some
Strategies for Low Drug Solubility 351

Fig. 14. The dissolution and precipitation of salts in the fasted stomach and during transit into the more neutral small intestine. (A) Dissolution of
a salt of a weak acid in the acidic stomach is promoted by the self-buffering effect of salts. Drug in solution in the stomach may enter the intestine and
be absorbed, or it may convert to the un-ionized free acid in the stomach, typically resulting in supersaturation and increasing the risk of precipitation.
Precipitated free acid may exist as fine particles (either amorphous or crystalline) or may form a layer of poorly soluble free acid around the salt
particle, inhibiting further gastric dissolution. On gastric emptying, dissolution of precipitated free acid is subsequently promoted by the higher pH in
the intestine. Incomplete dissolution of the salt of a weak acid in the stomach leads to gastric emptying of undissolved salt and the expectation of
improved dissolution in the higher pH environment of the small intestine. (B) Dissolution of the salt of a weak base in the stomach is favored by the low
pH environment. Drug in solution subsequently enters the intestine, where the increase in pH commonly leads to transient supersaturation as the
degree of ionization (and therefore drug solubility) is decreased. Depending on the rate of absorption, supersaturated drug may be absorbed or
precipitate (in either the amorphous or crystalline form). Precipitated free base may form discrete particles or coat incompletely dissolved salt particles,
inhibiting further dissolution. Redissolution is possible, however, as drug is absorbed and sink conditions return. Incomplete dissolution of the salt of
a weak base in the stomach results in gastric emptying of the undissolved salt, the dissolution of which is slow in the small intestine but assisted by the
self-buffering effect of the salt. Salt 5 light shaded purple, free acid/base 5 dark shaded purple.
352 Williams et al.

Fig. 15. The effect of acidity or basicity of weak acids and weak bases on the relationship between pHmax and physiologic pH, and therefore, the
likelihood of in situ salt-un-ionized drug conversion. In situ conversion from the more soluble salt to the less soluble free acid or free base is expected
where physiologic pH is lower (for weak acids) or higher (weak bases) than pHmax. For salts of typical weak bases (pKa ;9) under gastric and most
intestinal conditions, conversion to the free base is relatively unlikely. For weaker bases, however (pKa ;6), the likelihood of conversion increases,
especially under intestinal pH. For weak acids, conversion is more likely since GI pH values are largely acidic. For typical weak acids (pKa ;4), the
potential for conversion to the free acid is evident under gastric conditions and potentially in the small intestine (depending on pHmax). For even
weaker acids (pKa ;7) conversion under most physiologic conditions is likely. Note the difference between pHmax and pKa has been arbitrarily shown
here as ;2.5 pH units to allow simple illustration but will vary for differing salts.

cases, conversion to the hydrate may also occur during of dissolution of the anhydrate (Zimmermann et al.,
the early stages of dissolution, in turn reducing the 2009). Differences in the rates of crystallization of the
dissolution rate. Finally, where crystallization from free base were attributed to crystals of the anhydrate
supersaturated solutions occurs (in response, e.g., to acting as more effective nucleation substrates com-
self-buffering of a salt, gastric emptying of a weak base, pared with the hydrate. In contrast, under more acidic
or rapid dissolution of the anhydrous form), recrystal- conditions (i.e., ,pHmax), where conversion to the free
lization as the less soluble hydrate is also possible. The base was not anticipated, little difference in the
hydrate and anhydrous forms of a particular salt may dissolution behavior of the hydrate or anhydrous
show different propensities for conversion to the free material was evident (Zimmermann et al., 2009).
acid or free base, resulting in seemingly unusual Unfortunately, differences in sensitivity to conversion
differences in dissolution. For example, the pHmax of to the free base or free acid are seemingly unpredict-
siramesine hydrochloride is 4.8 for the anhydrate and able; in contrast to the data obtained with siramesine,
5.1 for the monohydrate (the higher pHmax for the diclofenac sodium trihydrate has been shown to
hydrate form likely reflecting the lower intrinsic convert to diclofenac acid more quickly than the
solubility of the drug; see Fig. 8). Dissolution of both anhydrous salt (Bartolomei et al., 2006).
the anhydrous and monohydrate forms at pH 6.4 d. Effect of Common Ions on Dissolution in the
(i.e., above the pHmax) therefore predicates the forma- Gastrointestinal Tract. As described in the previous
tion of a supersaturated solution and the likelihood of section, the presence of common ions reduces salt
precipitation of the free base. In this case, recrystalli- solubility. In light of the relationship between solubil-
zation of the free base occurred more readily from ity and dissolution rate, common ions may therefore
solutions of the anhydrous salt compared with the also suppress dissolution rate. For example, the
hydrate, with the surprising outcome of a slower rate dissolution of tetracycline hydrochloride in 0.1 M
Strategies for Low Drug Solubility 353

hydrochloric acid is slower than that of the free base, weak acids. Metal alkali salts tend to be more
and as a result, the oral bioavailability of tetracycline hygroscopic than organic amine salts; however, metal
hydrochloride is reduced compared with that of the free salts are often more soluble (Schwartz and Buckwalter,
base (Miyazaki et al., 1975, 1981). The addition of 1962; Hirsch et al., 1978). Increases in solubility and
sodium chloride to a dissolution medium can similarly dissolution rate through salt formation must therefore
reduce the dissolution rate of sodium salts (Mooney be considered in light of the stability properties of the
et al., 1981; Serajuddin et al., 1987). Although different new solid. For hygroscopic materials, interaction with
counterions may be sought to circumvent the effect of water vapor on storage may also result in hydrate
common ions, it is apparent that the selection of an formation, with the attendant risks of a reduction in
alternative counterion does not necessarily preclude solubility and dissolution rate (see Section IV.D.1.c).
a common ion effect since there is potential for in situ The transformation of a salt to a hydrate or other
conversion provided sufficient concentrations of the physical form on storage or hydration is often detri-
alternate counterion are present (Li et al., 2005a). mental to performance (i.e., aqueous solubility and
2. Physical Properties of Pharmaceutical Salts. dissolution rate) and highly dependent on choice of
Salts can exist in many chemical forms, including 1) counterion. However, such transformations are hard to
highly ordered crystalline phases, 2) crystalline solvates predict a priori, and as a consequence, the optimal salt
and hydrates (“pseudopolymorphs”), 3) liquid crystals, form is normally chosen after a salt screen (Gould,
and 4) irregular amorphous states. The propensity for 1986; Tong and Whitesell, 1998; Bastin et al., 2000).
a salt to exist in the crystalline or noncrystalline Pharmaceutical solids with low melting temper-
amorphous form is dependent on the properties of the atures are typically “soft and plastic,” whereas those
drug molecule, the counterion used, and also the solvent with higher melting temperature show “hard and
composition used to isolate the salt (Tong and Zografi, brittle” characteristics (Corrigan, 2007). Low melting,
1999; Giron and Grant, 2002; Black et al., 2007; Towler soft, and plastic materials typically process less readily
et al., 2008; Brittain, 2010). Crystalline salts are into free-flowing powders with good behavior on
generally favored for development over noncrystalline deformation (during compression), and this is often
forms owing to their superior stability and higher detrimental to the characteristics of the final dosage
melting temperatures. Salts formed using strongly form, such as disintegration, friability, and content
acidic or basic counterions typically show a greater uniformity. Salt formation usually leads to increases in
inclination to higher crystallinity (Hirsch et al., 1978; melting temperature compared with the free acid or
Bastin et al., 2000; Remenar et al., 2003a; Black et al., free base and therefore has the potential to enhance
2007; Gross et al., 2007). Amorphous salts of various processability. However, differences in packing effi-
drugs have been isolated (Tong and Zografi, 1999, 2001; ciency, and therefore differences in melting point
Towler et al., 2008; Hasa et al., 2011); however, these (Gould, 1986; O’Connor and Corrigan, 2001), also alter
are rarely pursued unless a formulation strategy that solubility, and as a result, there is some temptation to
subsequently stabilizes the amorphous form (such as select the salt with the lowest melting temperature as
a solid dispersion; see Section X) is envisaged (Towler this salt is likely to show the highest solubility (see
et al., 2008; Agoram et al., 2010). Section IV.A.4). Indeed, salt solubility commonly
Salts have also been widely explored as a means to decreases by a factor of 10 for every increase of 100°C
improve the chemical stability of weak acids (Walton in the melting temperature (Corrigan, 2007). A
et al., 1970; Cotton et al., 1994) and weak bases compromise is usually required between lower melting
(Bartolini et al., 1981; Walkling et al., 1983; Gould, points to encourage solubility and higher melting
1986; Lin, 1999; Ki et al., 2008). The stability benefits points to permit adequate processing. Salts with
accrued by isolation as a salt can be reversed, however, melting points below 100°C are rarely progressed
if the salt shows increased evidence of hygroscopicity (Stahl and Wermuth, 2002).
since moisture absorption can lead to reduced chemical 3. Potential Toxicity of Pharmaceutical Counterions.
stability and poorer flow and compression properties of The potential for toxicity has limited the exploration of
the solid drug (Gordon and Chowhan, 1987, 1990; many novel counterions; however, the wide range in
Adam et al., 2000). Salts absorb moisture by forming current use is usually sufficient for most applications.
hydrogen bonds with water in a mechanism analogous Of the counterions that have been previously used,
to interactions with water molecules in solution. Salts lithium salts are now generally avoided because of
that tend to show greater hygroscopicity include those their known pharmacological activities (Berge et al.,
containing more hydrophilic counterions, ostensibly as 1977), and bromide ions are also infrequently used
a result of the higher polarity of the crystal surface owing to their toxic effects and potential to accumulate
(Berge et al., 1977; Visalakshi et al., 2005; Gross et al., in the kidney (Torosian et al., 1973). Calcium salts
2007; Guerrieri et al., 2010). More hydrophilic counter- affect renal function (Schou, 1958; Berge et al., 1977),
ions include hydrochloride, dihydrochloride, and sul- and maleic acid salts cause renal tubular lesions in the
fate for weak bases and sodium and potassium for dog (Everett et al., 1993), although both are still widely
354 Williams et al.

used (Paulekuhn et al., 2007). Nitrates are also known exposure in vivo and facilitate more confident pro-
to cause GI irritation leading to nausea and vomiting gression through pharmacokinetic and toxicokinetic
(Berge et al., 1977). More recently, safety concerns studies. However, selecting the optimal salt is not
have been voiced surrounding the use of sulfonic acid straightforward, as many of the properties that dictate
counterions (mesylates, besylates, and tosylates). the potential in vivo performance of a salt (e.g., Ksp, the
These counterions are strongly acidic (pKa values potential for in situ conversions to a less soluble salt,
between 21.34 and 0.7; see Table 8) and consequently the formation of hydrates, self-association properties,
are widely used to form salts with weak bases (Engel and so forth) cannot be predicted using the fundamen-
et al., 2000; Gross et al., 2007; Kumar et al., 2007; tal solubility principles of the ionized form (i.e., via the
Elder et al., 2010). The antiviral product Viracept, for Henderson-Hasselbalch equation, eq. 14). The large
example, contains nelfinavir as a mesylate salt; number of potential counterions (Table 8) also precludes
however, its use was suspended by the European the synthesis and detailed evaluation of all potential
Medicines Agency in 2007 over concerns that it salt forms. A more efficient approach, therefore, is to
contained high levels of the sulfonate ester, ethyl develop experimental methodologies that rapidly screen
methanesulfonate (EMS) (Elder et al., 2010), which is potential salt forms that meet predetermined criterion.
genotoxic and possibly carcinogenic (Snodin, 2006). Shanker et al. (1994) described one of the first screening
More recent evidence, however, suggests that the high protocols to search for the most promising salt forms for
levels of EMS in Viracept resulted from contamina- new chemical entities, and several companies sub-
tion of the starting material rather than being a side sequently adopted similar screening approaches, with
reaction during salt formulation (EMEA, 2007). The each company setting its own selection criterion. The
use of equimolar concentrations of drug and counter- successes of salt screening methodologies have spawned
ion can eliminate the potential for EMS to form an era of automation and the creation of specialty
during salt manufacture (Teasdale et al., 2009; Elder companies dedicated to salt and polymorph screening.
et al., 2010), and mesylates remain popular as potential The development of small-scale characterization
counterions. techniques that require low (milligram) quantities of
In an attempt to simplify counterion choice based on raw material (Shanker et al., 1994; Tong and Zografi,
toxicity, counterions for oral or parenteral delivery 1999; Bastin et al., 2000; Morissette et al., 2004) and
have been categorized into three classes in accordance the availability of high-throughput screening methods
with their GRAS (“Generally Recognized as Safe”) for multiple physical forms now enable physicochem-
status and Accepted Daily Intake values. The classifi- ical profiling of salt forms to be a routine process that is
cation is summarized in Table 11, and more details on undertaken earlier in drug development (Kerns, 2001;
regulatory guidance for novel salt-formers and new salt Kerns and Di, 2003). One such example suggests that
forms for a previously approved drug are available in as little as 100 mg of drug material may be sufficient to
Verbeeck et al. (2006) and Stahl and Wermuth (2002). complete pharmaceutical salt profiling (Balbach and
Korn, 2004).
E. Feasibility of Salt Formation and Salt 1. Salt Formation Feasibility. A simple but effective
Selection Strategies indicator of the potential for salt formation is the
Early identification of the final chemical form for presence of a 2-unit difference in pKa between the drug
a potential drug product is of significant benefit to and the oppositely charged counterion (Tong and
rapid development. For poorly water-soluble drugs in Whitesell, 1998; Pudipeddi et al., 2002; Black et al.,
particular, finalizing a potential salt form early in the 2007). This stems from the realization that the wider
discovery program is expected to aid improved drug the difference in acidity and basicity between the drug

TABLE 11
“Criteria” of salt formers in classes 1, 2, and 3
Example
Classa Criteria
Acids Bases

1 Salt formers that can be used without restriction because Acetic, citric, fumaric, maleic, L-Arginine,
calcium, lysine,
they contain physiologically ubiquitous ions and/or ions hydrochloric, sulphuric, succinic magnesium, sodium, potassium
that occur as intermediate metabolites in biochemical
pathways. Frequently used in the past and present
2 Salt formers that while not naturally occurring have Besylate, mesylate, napsylate Diethylamine, tromethamine,
through a number of applications shown low toxicity nicotinate, tosylate,
and good tolerability
3 Salt formers that are occasionally used, mainly for the Nitric, formic, hydrobromide Piperazine, ethylenediamine
purposes of achieving ion-pair formation. Sometimes
suitable to solve particular problems
Data taken from Stahl and Wermuth (2002).
a
The different classes are designed to categorize salt counterions according GRAS (“generally recognized as safe”) status and Accepted Daily Intake values.
Strategies for Low Drug Solubility 355

and the counterion, the greater the probability of capacity to predict salt properties, salt selection
proton transfer during the acid-base reaction. Conse- remains largely empirical, and salt-screening techni-
quently, stronger counterions are often required for ques remain the mainstay of salt selection protocols.
salt formation using weakly acidic/basic drugs. For The major advances in this area have been the
example, phenytoin is a very weak acid (pKa 5 ;8–9), development of faster high-throughput processes that
and hence sodium (a strongly basic counterion) salts require less compound. Compared with traditional salt
are used in nearly all commercial products (exceptions synthesis, a high-throughput method may require only
include pediatric formulations and oral suspensions). milligrams (1–10 mg) of free acid/base per test, and
Notably, however, even the sodium salt is prone to with integrated automated or semiautomated processes,
conversion to the free acid during dissolution. Itracona- this can significantly reduce the number of steps required
zole is a very weak diprotic base (pKa,1 5 3.7, pKa,2 5 2), to isolate the salt form and increase the accuracy with
and documented exploration of the potential to form which this is performed. Strategies for the preparation
salts is therefore limited. In a small salt-screening of pharmaceutical salts have been recently reviewed
study, however, itraconazole reportedly formed several (Morissette et al., 2004; Kumar et al., 2007).
crystalline “hits” with a variety of different acid High-throughput salt screens provide an opportunity
counterions, including phosphoric acid, sulfonic acid, to assess rapidly a number of different counterions and
methanesulfonic acid, formic acid, and succinic acid. In to examine several experimental conditions that may
this study, however, the pKa values of the fumarate (pKa affect salt formulation (e.g., solvent type and composi-
5 3.03) and succinate (pKa 5 4.2) counterions resulted tion). For example, Ware and Lu (2004) describe
in differential pKa values that did not meet the “rule of a method to evaluate 104 potential salts of a candidate
2,” and the crystalline materials recovered may have weak base using 13 potential counterions and 8 dif-
been itraconazole salts, but they may also have been ferent solvents. In the first stage, 100 mL of a solution
hydrates or cocrystals (Tarsa et al., 2010). containing 5 mg of drug (in this case, trazadone free
When attempting to develop salts of very weak bases, base) was robotically dispensed into individual wells of
the potential for reversion to the free base (dispropor- a 96-well plate, followed by 200 ml solvent and 15 ml of
tionation) in the solid state must be addressed. Dispro- acidic counterion (in an equimolar ratio). The 96-well
portionation can occur when basic excipients increase plates were incubated at ambient temperature for 12 h,
the microenvironmental pH in the dosage form above during which time the solvent evaporated and salts (if
the pHmax of the salt, resulting in the loss of a proton formed) precipitated or crystallized. Samples of solid
(Guerrieri and Taylor, 2009). Factors that increase the material from each well were subsequently analyzed
risk of disproportionation include weak basicity (low pKa using polarized-light microscopy for crystallinity. In
values), high salt-base solubility ratios, and low pHmax stage 2, 25-mg batches of 34 potential “hits” from the
(Guerrieri and Taylor, 2009). Disproportionation can lead first stage were characterized by XRPD and DSC before
to decreased rates of dissolution (Rohrs et al., 1999), evaluation of hygroscopicity and melting point. Finally,
decreased potency (Zannou et al., 2007), and changes to 1-g batches of four salts were prepared, and the particle
tablet hardness and disintegration (Williams et al., 2004). size, surface area, and solubility were determined (Ware
Additional complexity also surrounds the formation and Lu, 2004).
of salts of diprotic drug molecules. For a dibasic High-throughput salt-screening methods similar to
compound, for example, avitriptan (pKa,1 5 8.0, pKa,2 those described by Ware and Lu allow a large number
5 3.6), decreasing pH results in the protonation of the of drug-counterion combinations to be randomly screened
most basic moiety and subsequent formation of the for potential salt hits. However, smaller, more directed
monosalt on reaching the first (highest) pHmax. Further screens are possible if the counterions tested are chosen
decreases in pH protonate the second (less basic) only from those that are likely to provide a shift in pH
moiety and eventually the formation of the disalt at sufficient to shift the pH above (for weak acids) or below
pH values below the second pHmax (‘pHmax,2’). To form (for weak bases) the corresponding pHmax (Serajuddin
the disalt, the counterion must therefore be sufficiently and Pudipeddi, 2002; Serajuddin, 2007). This approach
acidic to protonate both basic groups. In the case of has the potential to eliminate several potential counter-
avitriptan, the hydrochloride anion was sufficiently ions by virtue of an inappropriate pKa.
acidic to form the disalt while mesylate, acetate, lactate, It is also apparent that general screening methods
tartrate, and succinate were all unable to lower the pH and selection criteria might not be appropriate for
below pH 5 and therefore formed only monosalts drugs where solubility is a critical limitation to oral
(Serajuddin and Pudipeddi, 2002). Similarly, dibasic bioavailability since salts that provide significant solu-
quetiapine (pKa,1 5 6.83, pKa,2 5 3.56) formed disalts bility advantage might be discarded early in a general
with hydrochloride, whereas, with fumaric acid, only screen because of “failing” progression criteria for other
a monosalt was formed (Volgyi et al., 2010). properties such as hygroscopicity. For these compounds,
2. Salt-Screening Strategies. Because of the com- the selection of hygroscopic salts may be acceptable
plexities associated with salt formation and the limited where solubility properties are sufficiently advantageous
356 Williams et al.

and where hygroscopicity can be addressed through the state stability. Under conditions where bulk solution
design of suitable packaging materials or via a more pH is buffered, the solubility advantages of the pH shift
tailored choice of excipients to minimize water uptake on promoted by salt formation may be reduced; however,
storage (Kumar et al., 2008b). Solubility-focused screens the dissolution rate often remains high as a result of
have therefore been proposed (Huang and Tong, 2004) a “self-buffering effect” where the microenvironment
where an in situ measurement of salt solubility is chosen pH in the unstirred water layer surrounding a dissolv-
as the primary screening outcome. Solubility-screening ing salt form promotes drug ionization and dissolution.
methods have been described by Shankar et al. (1994) Subsequent diffusion of dissolved drug into the bulk
and later by Tong and Whitesell (1998), where the pH may result in precipitation, but where the rate of
primary objective of the screen was to select a salt based absorption is faster than the rate of precipitation, or
on solubility improvements sufficient to support de- where supersaturation is stabilized, significant increases
velopment. The counterions investigated (hydrochloride, in absorption remain possible. Several examples high-
mesylate, phosphate, citrate, succinate, or tartrate) were lighting the ability of salts to improve oral bioavailability
chosen based on pKa values that were more than two are evident in the literature, and the abundance of salt
units below the pKa of the drug to ensure a pH shift forms in marketed products is tribute to the potential
sufficient to lower pH below the pHmax and to provide advantages that may accrue as a result of isolation as the
greater assurance that any isolated solid was a salt. In salt form.
each case, a counterion concentration of 0.1 M was However, prediction of the solubility benefits of
sufficient to ensure that the counterion was in excess different salt forms remains difficult, and salt-screening
(relative to the quantity of drug) and, accordingly, that techniques provide the mainstay industrial approach
the quantity of dissolved drug was dependent on the for salt selection. Salts may also suffer from reductions
solubility of the formed salt. in apparent solubility because of the presence of ions in
Several other strategies for salt selection have been the dissolution media (in vitro or in vivo) that are com-
described (Gould, 1986; Morris et al., 1994; Bastin mon with the counterion used for salt formation (the
et al., 2000) and reviewed (Corrigan, 2007; Serajuddin, “common ion effect”) or as a result of in situ conversion
2007; Kumar et al., 2008b), and the interested reader is to salts with lower solubility to the free acid or free base
directed to these references for more detail. or to a hydrate (Table 12). The presence of ionizable
F. Summary groups also does not guarantee the formation of a soluble
and stable salt, and drug and counterion should be
Many new chemical entities contain functional matched to provide differences in pKa of at least 2 units
groups that are ionizable under physiologic conditions. to ensure proton transfer.
In general, the ionized form has greater aqueous Assessment of the feasibility of salt isolation and the
solubility than the un-ionized species; therefore, choice of counterion are therefore critical determinants
methods that promote ionization typically enhance of the success of pharmaceutical salts. However, for
aqueous solubility. This can be achieved in solution ionizable compounds, the relative simplicity of phar-
dosage forms by modification of solution pH (commonly maceutical salts, the potential benefits in dissolution
via the use of buffers) and in the case of solid dosage rate and solubility, and the presence of a significant
forms via the isolation of salts. industrial experience base dictate that the isolation of
For parenteral formulations, an increase in drug
solubility via simple pH adjustment can be sufficient
TABLE 12
for complete solvation of the target drug dose, and as Summary of factors affecting the formation, solubility and dissolution
a rule of thumb, differences in pH of 1 unit away from rate of salts
the pKa of the drug result in 10-fold enhancements in Key factors that affect salt formation
solubility. In the presence of buffers or other counterions, pKa difference between drug and available
counterion
however, the upper solubility limit is determined by the Drug solubility
solubility product of the salt that is formed in situ. Salt solubility
Buffer choice and appreciation of the potential counter- Organic solvents
Key factors that affect salt solubility
ions that might be present in, for example, plasma, are Combined energies of solvation of drug and
therefore important determinants of practical solubility counterion
Lattice energies
limits. Where pH adjustment alone is insufficient to Common ions
allow appropriate solubilization, pH adjustment may be Key factors that affect salt dissolution rate
Salt solubility
used in combination with other strategies, including Self-buffering capacity
cyclodextrins, cosolvents, and nonionic surfactants. Environmental pH of the surrounding fluids
Isolation of a salt and the subsequent dissolution of Common ions
In situ conversion from salt to free acid/free base
the salt in vivo provide many of the advantages of pH Hydrate formation during storage or during
adjustment but also allow development of a solid dissolution
In situ interconversion between salt forms
dosage form, with the attendant advantages in solid-
Strategies for Low Drug Solubility 357

a preferred salt form remains a priority when attempt- polymorphs or the amorphous form exhibit different
ing to address the problems of low solubility of free energies, which in turn give rise to unique
a potential drug candidate. physicochemical properties. In the context of the current
review, different crystalline polymorphs (or the amor-
phous form) have different solubilities and, therefore,
V. Optimization of Crystal Habit: Polymorphism
have the potential to impact oral bioavailability for
and Cocrystal Formation
drugs where low aqueous solubility is a significant
Crystal engineering is traditionally defined as the limitation to absorption (Aguiar and Zelmer, 1969; Law
deliberate design and control of molecular packing et al., 2004).
within a crystal structure with the intention of Cocrystals constitute a molecular complex between
generating a solid that shows a particular and desir- a drug and cocrystal former and also result in changes
able characteristic (Desiraju, 2001, 2010). Since the to the crystal lattice. In some respects, cocrystals are
solid-state properties of a drug will influence a range of analogous to pharmaceutical salts in that they consti-
different drug properties, including solubility, the level tute a complex between drug and an additional species
of interest in crystal engineering strategies has in- (counterion, coformer). Unlike the salt complex formed
creased dramatically over recent years (Desiraju, 2001; between a weak acid or weak base and the respective
Datta and Grant, 2004; Blagden et al., 2007). counterion, in the case of a complex between a drug
In the broadest sense, crystal engineering may and a cocrystal former, proton exchange does not occur
encompass any manipulation that results in altered (Aakeroy et al., 2007). Unlike salts, cocrystallization
crystal packing, including both traditional approaches, strategies are therefore not limited to ionizable
such as the isolation of different salts (Peterson et al., compounds.
2010), solvates (Blagden et al., 2007), or polymorphs, Isolation of a compound as a high-energy crystalline
and more recent approaches, such as cocrystal forma- polymorph or in the amorphous form can have a pro-
tion (Aakeroy and Salmon, 2005). In each of these found impact on apparent drug solubility. Indeed, on
approaches, a change to solid-state properties is likely dissolution, transient differences in the amorphous-to-
to affect solubility and may be advantageous in crystalline drug solubility ratio may cover several
overcoming solubility limitations that stem from the orders of magnitude, leading to potentially dramatic
strength of the crystal lattice. In the case of salt increases in dissolution rate. These solutions are,
formation, however, changes to solubility reflect both however, thermodynamically unstable, and over time,
changes to the crystal lattice and coincident changes to drug in solution will recrystallize, reverting to more
pH and ionization (and where changes to pH and thermodynamically stable (and generally less soluble)
ionization provide the primary beneficial effects). Salt forms (Hancock and Zografi, 1997; Murdande et al.,
formation is therefore addressed in a separate section 2011a). These may include, for example, more stable
(see Section IV). crystalline polymorphs or hydrates (Pudipeddi and
Here the discussion of crystal engineering strategies Serajuddin, 2005; Agoram et al., 2010; Murdande
is limited to polymorphism and cocrystal formation et al., 2011a), but in both cases, solubility is usually
(i.e., approaches that alter the crystal habit). A reduced. This transformation can occur very rapidly on
somewhat broader use of the term crystal engineering dissolution, complicating solubility measurements for
has also emerged (Biradha et al., 2011) to describe the high energy form, and can also occur during storage
strategies that include not only deliberate manipula- in the solid state, particularly on exposure to moisture
tion of the crystal lattice but also manipulation of or heat (Yu, 2001; Pudipeddi and Serajuddin, 2005;
crystallization conditions to isolate drug crystals of Janssens and Van den Mooter, 2009). A well known
a particular size and shape (Nokhodchi et al., 2003; example of the complexities of polymorphic transitions
Blagden et al., 2007). The latter strategy, however, is is that of ritonavir (Norvir; Abbott Laboratories, North
arguably better described as particle (rather than Chicago, IL), which was initially launched by Abbott
crystal) engineering since it is the physical properties Laboratories in 1992 as a capsule formulation contain-
of the particle that are changed rather than the crystal ing dissolved drug at a loading that was believed to be
form (see Section IX). Indeed, the same polymorphic below the solubility of the thermodynamically stable
form can form particles with widely differing character- crystal form (to avoid the potential for drug crystalli-
istics (Nokhodchi et al., 2003; Iacocca et al., 2010). zation during storage). However, a previously unidenti-
The potential for a drug to form one or more fied polymorph of enhanced stability, but correspondingly
crystalline solids that differ only by the molecular lower solubility, was subsequently identified during
arrangement of drug molecules in the crystal lattice is a change to process chemistry. The lower solubility of
referred to as polymorphism, and different crystalline the more stable form drew into question the long-term
forms of the same compound are described as poly- stability of the original formulation, prompting product
morphs. Materials lacking crystalline character are recall and reformulation prior to rerelease (Chemburkar
said to be amorphous. Since they differ in crystal habit, et al., 2000; Bauer et al., 2001; Huang and Tong, 2004).
358 Williams et al.

Deliberate isolation of high-energy crystal forms for the molecules in an orientation that is constant through-
purpose of improved solubility is therefore possible only out the crystal lattice. The nature and strength of the
when coupled with formulation strategies that impart intermolecular interactions that form between drug
a level of stabilization to the crystal form sufficient to molecules within the unit cell and between adjacent
cover the required shelf-life. In this regard, recent cells are determined by the chemical properties of the
increases in the understanding of crystal transitions drug molecule and their orientation within the unit cell.
and the widespread availability of thermal analytical Molecular interactions within a crystal lattice are
technologies have allowed a more rational approach to either attractive or repulsive and include relatively
developing formulations that are stable over timescales weak van der Waals forces, p-p stacking, and
relevant to the typical shelf-life of a drug product electrostatic interactions (0.522 kJ per mol), stronger
(Hancock and Zografi, 1997; Weuts et al., 2011), and hydrogen bond interactions (5–30 kJ per mol), and much
coformulation with (usually polymeric) materials that stronger (;150 kJ per mol) ionic interactions. The
are capable of stabilizing drug in a high-energy form summation of the energies from all intermolecular
over long periods is becoming routine (Taylor and interactions determines the total packing energy of
Zografi, 1997; Friesen et al., 2008; Curatolo et al., a crystal or the crystal lattice energy. Since stronger
2009). The most common of these approaches is the intermolecular interactions minimize the molecular
formulation of polymorphic or amorphous drug in energy (and mobility) of the molecules within the
polymeric solid dispersions, and this formulation tech- crystal, the higher the lattice energy, the greater the
nology is described in more detail in Section X. stability (and lower the solubility) of the crystal (Datta
In the current section, we introduce the fundamental and Grant, 2004; Lohani and Grant, 2006). The most
concepts of crystallinity and polymorphism in pharma- stable crystal form therefore allows the largest number
ceutical solids and explore the relationship between of molecular interactions with adjacent molecules.
the solid-state properties of a drug and solubility. The Accordingly, a crystal lattice in which the orientation
latter provides the rationale for the isolation of high- of the drug molecule does not permit the maximum
energy polymorphs and also serves as background to number of interactions with adjacent molecules is
subsequent sections that discuss formulation approaches unstable, or metastable, and over time or after exposure
that may be used to stabilize high-energy solids (usually to heat or moisture, molecules within the crystal lattice
the amorphous form) such as solid dispersions. The will reorientate (relax) to adopt a more thermodynam-
concept of crystal engineering via nontraditional means ically stable conformation. It is important to recognize
is subsequently introduced and, in particular, the design the potential confusion that can occur between reference
and isolation of cocrystals. Cocrystals provide the to unstable polymorphs as “high-energy” solids (where
potential opportunity to match the benefits of high- the reference to high energy refers to thermodynamic
energy polymorphs or amorphous drug, such as improved instability) and the realization that the crystal lattice
solubility and dissolution characteristics, with the de- energy of these systems is, in fact, low compared with
sired stability traits of a stable crystal form (Blagden the thermodynamically stable crystal form.
et al., 2007; Desiraju, 2010). Polymorphism is the term most frequently used to
describe the ability of an organic molecule to exist in
A. Polymorphs more than one crystalline form (i.e., polymorph) (see
1. Crystal Packing, Polymorphism, and Phase Trans- Fig. 16), and it has been estimated that approximately
formations. In an ideal crystal, identical structural 80% of pharmaceutical compounds show more than one
components, or unit cells, are repeated regularly and polymorph (Grunenberg et al., 1996; Datta and Grant,
indefinitely in three dimensions in the absence of 2004; Lohani and Grant, 2006). In addition to “true”
imperfections or defects (Vippagunta et al., 2001). In polymorphs, where different crystal packing arrange-
pharmaceutical crystals, the unit cell comprises drug ments result entirely from differences in packing

Fig. 16. Schematic diagram illustrating common solid-state arrangements of drug molecules in the crystalline state and the amorphous form. Drug
polymorphs vary only in the molecular organization of individual drug molecules in the crystalline state. Solvates and hydrates (often referred to as
pseudopolymorphs) are molecular complexes of drug and molecules of water (in the case of hydrates) or other solvent molecules (in the case of solvates)
that form intermolecular interactions with the drug and become incorporated within the crystalline lattice. Cocrystals are also molecular complexes,
consisting of drug and a cocrystal former (coformer), which interact primarily via hydrogen bonds. Popular definitions of cocrystals in the
pharmaceutical literature state that the coformer must be solid at room temperature, distinguishing cocrystals from solvates and clathrates. Salts are
in many ways analogous to cocrystals but differ by virtue of an electrostatic interaction that forms between a charged drug molecule and an
oppositively charged counterion.
Strategies for Low Drug Solubility 359

efficiency of drug alone, alternate crystal structures


may also form following a disturbance to the crystal
packing arrangement induced by the presence of an
additional molecule. For example, where molecules
favor hydrogen bond formation with water or other
polar solvents, the solvent molecules may be incorpo-
rated into the host lattice during crystallization,
forming stoichiometric or nonstoichiometric hydrates
(where the solvent is water), or solvates (for other
solvents). The presence of solvent molecules within the
crystal lattice alters the pattern of drug-drug inter-
actions and, in doing so, gives rise to altered crystal
arrangements. However, solvates and hydrates are not
polymorphs in the true sense and, thus, are often
referred to as pseudopolymorphs. Similarly, the amor-
phous state should not, in theory, be described as
a different polymorphic form as this state lacks long-
range order (Datta and Grant, 2004) and can demon-
strate only a certain degree of local order, for example,
symmetrical carboxylic dimerization in the case of
indomethacin molecules (Taylor and Zografi, 1997;
Rodriguez-Spong et al., 2004). Pragmatically, however,
the US Food and Drug Administration (FDA) and the
International Conference on Harmonization (ICH)
classify hydrates, solvates, and amorphous forms all
as polymorphs (Ku, 2010).
2. Effect of Polymorphism on Drug Solubility, Disso-
lution Rate, and Oral Absorption. Recent analysis of
the effect of polymorphism on the solubility and dis-
solution rate of 55 compounds suggested that the dif-
ferences in apparent solubility between a high-energy
crystalline polymorph and the most stable crystal form
are, in most cases, relatively moderate. In this analysis, Fig. 17. (A) Kinetic solubility of selected rifaximin polymorphs in water;
the solubility ratio between high-energy crystal forms (B) mean concentration-time profiles of rifaximin polymorphs after oral
administration of 100 mg/kg to four dogs. Adapted from Viscomi et al.
and the most stable crystalline polymorphs was be- (2008).
tween 2 and 3 for most of the compounds and less than 5
for all but one compound (Pudipeddi and Serajuddin,
2005). In contrast, the amorphous:crystalline solu- energy crystalline polymorphs can lead to significant
bility ratio is typically much higher (Hancock and increases in oral bioavailability for drugs including
Parks, 2000; Murdande et al., 2010b). Solubility com- chloramphenicol palmitate (Aguiar et al., 1967), sulfa-
parisons have also been made between hydrates and meter (Khalafallah et al., 1974), phenylbutazone
anhydrous crystalline forms across a collection of 17 (Pandit et al., 1984), amobarbital (Kato and Kohketsu,
compounds, and only three compounds showed a 1981), and more recently celecoxib (Lu et al., 2006) and
greater than 3-fold difference in solubility between rifaximin (Viscomi et al., 2008) (Fig. 17, A and B). The
the hydrate and the anhydrous form (Pudipeddi and seemingly contradictory nature of in vitro solubility
Serajuddin, 2005). Although there are occasional ex- assessments for high-energy polymorphs and in vivo
ceptions, hydrates are generally less water-soluble observations may result from the inherent difficulties
compared with anhydrous forms of the same drug in assessing the true solubility advantage of a thermo-
(Huang and Tong, 2004), a situation reflecting the dynamically unstable solid in vitro. Thus, only one true
reduction in available sites for drug interaction with equilibrium solubility value is evident for a drug (the
water in the hydrate. solubility of the lowest energy, thermodynamically
On the basis of solubility alone, there would appear stable crystal form). In practice, equilibration to form
to be only moderate potential benefit (2- to 5-fold the thermodynamically stable crystal form may be
increase in solubility) for poorly water-soluble drugs in rapid or may occur over an extended time. It may be
using different crystalline polymorphic forms or by possible, therefore, to obtain a solubility value for the
switching between an anhydrous and hydrated form thermodynamically unstable solid, provided solubility
(Singhal and Curatolo, 2004). However, the use of high- measurements are made before transformation to the
360 Williams et al.

more thermodynamically stable species. In many following sections of this review that address the
cases, however, recrystallization to the lower-energy specific formulation approaches that might be used to
(more stable) polymorph is rapid, leading to underes- stabilize amorphous drug, in particular, the use of solid
timation of the apparent solubility of the high-energy dispersion technologies (see Section X).
polymorph. Nonetheless, the rate of dissolution of Nonetheless, amorphous drug has the potential to be
a high-energy polymorph can be many times faster used in preclinical settings to afford increased drug
than that of the equivalent lower-energy material, and exposure for efficacy and safety assessment and where
drug concentrations in aqueous solution that are long-term stability of the formulated drug candidate is
orders of magnitude higher than the solubility of the not essential (Nagapudi and Jona, 2008; Byrn et al.,
thermodynamically stable crystal may be obtained. 2010; Palucki et al., 2010; Zheng et al., 2011). Indeed, it
This transient increase in apparent solubility may be is possible that the materials emerging from early
described as the kinetic solubility of the high-energy chemistry programs (i.e., before formal polymorph and
polymorph, that is, the solubility that is maintained crystallization screens) may contain at least some
for a finite period (and that reflects the solid-state amorphous material, and this may provide assistance
properties of the high-energy polymorph rather than in early exposure studies. The corollary to this sugges-
the equilibrium solubility of the stable crystal) but is tion, however, is that the advent of stricter control of
effectively transient. Where the rate of recrystalliza- polymorphic form in subsequent drug batches may lead
tion of the stable crystal is slow or deliberately delayed, to a decrease in oral bioavailability for poorly water-
high initial drug concentrations may be maintained in soluble drugs as the amorphous content is reduced.
the GI tract for sufficient periods to drive significant Early identification of possible polymorphs and confir-
increases in drug absorption. When assessing the in mation of the stable crystal form is therefore important,
vitro solubility and dissolution behavior of different with the latter providing a stable comparative baseline
polymorphic or amorphous forms, data obtained over for subsequent bioavailability-enhancing approaches.
shorter (non-equilibrium) timescales may therefore Beyond effects on solubility, dissolution rate, and
provide a more accurate indication of in vivo dissolu- stability, different polymorphs also exhibit differences
tion (Singhal and Curatolo, 2004; Blagden et al., 2007). in a range of pharmaceutical properties (e.g., hygro-
In Fig. 17A, for example, different rifaximin crystalline scopicity, flow, compaction, and compatibility), all of
polymorphs show significant differences in kinetic which may be a critical in determining whether a drug
solubility, and these are reflected in differences in in candidate becomes a marketed product. These proper-
vivo oral absorption (Fig. 17B). In contrast, estimates ties, although important, lie outside the scope of the
closer to the equilibrium solubility of the various current review. Further information concerning the
polymorphs (e.g., the data in Fig. 17A after ;5 h) suggest physical properties of drug polymorphs and pseudopo-
much more moderate differences in drug solubility and lymorphs can be found in relevant reviews (Morris
values that correlate less effectively with the in vivo data. et al., 2001; Zhang et al., 2004).
The isolation and formulation of high-energy poly-
morphs (or amorphous drug), therefore, have the B. Cocrystals
potential to enhance significantly the bioavailability Cocrystals provide an alternative crystal engineer-
of poorly water-soluble drug molecules when compared ing strategy for improved drug solubility (Vishweshwar
with the most stable crystal form. It is also evident that et al., 2006; Henck and Byrn, 2007; Childs and
this is complicated by thermodynamic instability when Zaworotko, 2009; Friscic and Jones, 2010). Cocrystal
compared with the most stable crystal form, both in formation places greater emphasis on prospective
terms of the stability of the polymorph in the dosage molecular design of the crystal form, and has the
form and the generation of a supersaturated solution added advantage over the isolation of high-energy
on dissolution. In practice, the key to harnessing the crystal forms of superior thermodynamic stability in
utility of high-energy polymorphs lies in stabilizing the the solid-state (Fleischman et al., 2003; Good and
polymorph in the formulation and in stabilizing the Rodriguez-Hornedo, 2009).
supersaturated solution that forms in vivo. As such, 1. Cocrystals as a Mechanism of Enhanced Drug
high-energy polymorphs or amorphous drug are rarely Solubility and Dissolution. Pharmaceutical cocrystals
administered unformulated or as a simple suspension are defined as mixed crystals consisting of two or more
[with one notable exception being Zinnat tablets molecular species (that alone are solid at ambient
(GlaxoSmithKline), which contains amorphous cefur- conditions) held together by noncovalent and nonionic
oxime axetil] and instead are most commonly coformu- forces (Vishweshwar et al., 2006; Arora and Zaworotko,
lated with stabilizing agents such as polymers to form, 2009). Although there is considerable debate over the
for example, solid dispersion formulations. Broader definition of a cocrystal (Vishweshwar et al., 2006;
description of the application of polymorphs, or Blagden et al., 2007; Schultheiss and Newman, 2009),
amorphous drug, to the enhancement of drug dissolu- here we exclude crystal solvates, hydrates, and
tion and solubility is therefore also found in the clathrates as a component of each of these complexes
Strategies for Low Drug Solubility 361

will exist in either the liquid or gaseous state at solubility can be defined by the equilibrium constant
ambient temperature (Morissette et al., 2004; Aakeroy that describes cocrystal dissociation. Thus, for a 1:1
and Schultheiss, 2007). cocrystal of drug A and a coformer B in equilibrium with
Cocrystals are distinguished from salts by the pres- A and B in solution, the equilibrium relationship (eq. 25)
ence of neutral molecular species within the cocrystal and equilibrium constant or solubility product (Ksp in
lattice rather than a molecular complex containing drug eq. 26) can be described. Note that in this case, the
and counterion in a charged form (Fig. 16) (Aakeroy activity coefficients for A and B in solution are assumed
et al., 2007; Shan and Zaworotko, 2008; Friscic and to be unity and the activity of the solid crystal invariant—
Jones, 2010; Stevens et al., 2010). However, analogous to both assumptions that hold true for sparingly soluble
salts, cocrystals may provide dramatic increases in systems (Good and Rodriguez-Hornedo, 2009):
dissolution rate compared with simple drug crystals,
and this effect can be ascribed to an improvement in Ksp
A a Bb ⇌ A a 1 Bb ð25Þ
the drug solubility (Banerjee et al., 2005; Good and
Rodriguez-Hornedo, 2009).
K sp 5½Aa ½Bb ð26Þ
The molecular components of cocrystal complexes
are the target molecule (traditionally the drug mole- The solubility product Ksp is unique for a cocrystal in
cule) and the cocrystal former(s) (also known as the a particular solvent and describes the dissociation of
cocrystallizing agent or coformer). Cocrystals are formed the molecular complex in solution. Ksp is dependent on
by intermolecular interactions, or synthons, that form the strength of the solid-state interactions between the
between the drug and a potential cocrystal former, drug and cocrystal former relative to their interaction
which in turn leads to the creation of supramolecular with the solvent (Good and Rodriguez-Hornedo, 2009).
assemblies. Synthons common to pharmaceutical coc- From eq. 26, it follows that for a 1:1 drug-coformer
rystals are based largely on hydrogen bonding motifs complex (i.e., a 5 b 5 1), the solubility of the cocrystal
and are categorized either as heterosynthons (e.g., is equal to the square root of Ksp. It is also apparent
carboxylic acid-amide, pyridine-carboxylic acid, and that the solubility of the cocrystal is dependent on the
imidazole-carboxylic acid dimers) or as homosynthons
concentration of coformer in solution, such that
(e.g., carboxylic acid-carboxylic acid and amide-amide
increases in coformer concentration reduce the solubil-
dimers) (Fig. 18). Cocrystal formation is favored when
ity of the cocrystal in a similar fashion to the concept of
noncovalent intermolecular interactions between com-
the common ion effect for sparingly soluble electrolytes.
plementary functional groups on the drug and the
Depending on the choice of coformer, the solubility of
cocrystal former are more energetically favorable than
the cocrystal may be higher or lower than the
intramolecular interactions between two drug molecules
equilibrium solubility of drug alone. Cocrystals may
(Chow et al., 2008). Conversely, where drug-drug rather
be designed to address a number of pharmaceutical
than drug-coformer interactions dominate, recrystalli-
issues, but here we focus on the use of cocrystals to
zation of two distinct solid phases (i.e., drug and
promote solubility and dissolution rate, and under
coformer) is evident rather than cocrystallization to
form a single solid phase (Aakeroy and Salmon, 2005). these circumstances, coformers are chosen such that
Similar to interactions between a drug and its the solubility of the cocrystal is in excess of that of the
counterion in a sparingly soluble salt, cocrystal drug alone. In these instances, drug concentrations in
solution resulting from dissolution of the cocrystal will
be in excess of the solubility of the thermodynamically
stable drug crystal (Nehm et al., 2006; Good and
Rodríguez-Hornedo, 2009; Schultheiss and Newman,
2009). The system is therefore metastable, and the
solution is supersaturated with respect to drug solubility
in the absence of coformer. This is a similar situation to
that encountered on dissolution of a high-energy poly-
morph, where the initial solubility is high, but over
longer timescales, reversion to the solubility of the stable
crystal form will occur. In practice, however, significant
increases in dissolution and apparent solubility are
possible, depending on the rate of recrystallization from
the supersaturated solution (see Section V.B.5).
In contrast to most high-energy polymorphs, the
intermolecular interactions between drug and coformer
Fig. 18. Common supramolecular synthons that may form in a cocrystal in the solid state (which are a necessary part of co-
between the drug and a suitable cocrystal former (coformer). crystal formation) stabilize the cocrystal lattice such that
362 Williams et al.

issues surrounding solid-state stability are circumvented


(or at least reduced). Simplistically, cocrystals poten-
tially provide solubility benefits that are analogous to
those provided by high-energy crystalline polymorphs
or amorphous drug but with solid-state stability
advantages similar to that of salts.
2. Solubility Assessment of Cocrystals. As with
high-energy polymorphs, equilibrium solubility testing
of cocrystals is complex since the kinetic solubility of
the cocrystal (reflecting Ksp) is usually greater than
Fig. 19. Phase solubility diagram showing the effect of coformer
that of the equilibrium solubility of the stable drug concentration on the solubility of a cocrystal, shown as the solid red line.
crystal. Dissolution of a cocrystal, therefore, results in The dashed horizontal line represents the solubility of the drug in the
drug concentrations that are typically greater than the absence of coformer. Four regions of the phase solubility diagram are
identified: i) where excess solid material consists of the drug; ii) and iv)
solubility of the stable crystal, and drug precipitation is where excess solid material consists of the cocrystal; and iii) where all
thermodynamically favored until such time as the solid drug and cocrystal are dissolved. Arrows describe cocrystal pre-
cipitation (green arrow) and cocrystal dissolution (blue arrow) methods for
intrinsic solubility limit of the most thermodynami- determining the solubility of the cocrystal. The red line is described by [A]a
cally stable solid species (i.e., the most stable drug 5 Ksp/[B]b. Adapted from Good and Rodriguez-Hornedo (2009).
polymorph) is reached. Moreover, as cocrystals give
rise to increases in transient solubility, which in turn
may lead to significant increases in dissolution rate stable species—knowledge that greatly assists in the
and absorption, measurement of the equilibrium isolation of pure cocrystal (see Section V.B.4).
solubility is likely to underestimate the potential in The effect of increasing the coformer concentration
vivo advantages of cocrystals. Therefore, simple meth- on the solubility of a cocrystal is illustrated in the
ods to assess cocrystal solubility and to distinguish phase solubility diagram in Fig. 19, where the solid red
line depicts decreasing cocrystal solubility with in-
cocrystal solubility from the equilibrium solubility of
creasing coformer concentration and the horizontal
the system (which intrinsically moves toward the
dashed line represents the solubility of the drug crystal
solubility of the thermodynamically stable crystal form
in the absence of coformer. The solubility of the
of pure drug) are required.
coformer is assumed to be significantly in excess of
A simple measure of cocrystal solubility may be
that of drug and cocrystal and is therefore not shown.
gained via assessment of the kinetic solubility of the
Four regions are evident in the phase diagram. Region
complex using traditional phase solubility methods. In
1 is where excess solid phase comprises crystalline
this case, the assumption must be made that limited
drug, since the concentrations of drug and coformer in
recrystallization occurs from the metastable solution
solution are greater than the drug crystal solubility
that results on dissolution (Schultheiss and Newman,
limit but below the solubility of the cocrystal; region 2
2009). Alternatively, Good and Rodriguez-Hornedo is where concentrations in solution are in excess of both
(2009) recently described a method to determine more drug and cocrystal and excess solid phase consists of
accurately the solubility of a cocrystal based on an the species with the highest solubility (i.e., the cocrys-
understanding of cocrystal-phase solubility relation- tal); region 3 is where the solution is undersaturated
ships (Good and Rodriguez-Hornedo,. 2009) This ap- with respect to both drug and cocrystal (and therefore
proach is predicated on the realization that increasing no solid phase is evident); and region 4 is where
the concentration of coformer in solution decreases the cocrystal alone is the thermodynamically stable solid
solubility of the cocrystal (eq. 26) in a situation phase as concentrations are lower than the solubility of
analogous to the common ion effect with sparingly drug but greater than that of the cocrystal. At the
soluble salts. For cocrystals, where the cocrystal solu- transition point (labeled X), the two solid phases (drug
bility is higher than that of the parent drug, adding and cocrystal) coexist in equilibrium with a solution
coformer to the system reduces cocrystal solubility, containing dissolved drug and coformer (Good and
eventually leading to a point where the solubility of the Rodriguez-Hornedo, 2009). The transition point pro-
drug and that of the cocrystal are equal. This is an vides the coformer concentration at which the solubil-
invariant point on the phase solubility diagram, and ity of the drug is equal to the cocrystal solubility and,
measurement of drug and coformer solution concen- consequently, allows calculation of Ksp and ultimately
trations in equilibrium with both solid drug and solid the solubility of the cocrystal (Good and Rodriguez-
cocrystal at this transition point allows for calculation Hornedo, 2009).
of Ksp. From Ksp, the intrinsic solubility of the cocrystal Experimentally, the transition point concentrations of
can be calculated. A further advantage of this method drug and coformer can be determined using the reaction
is the identification of regions in the phase solubility crystallization method (RCM), in which drug crystals
diagram where the cocrystal is the thermodynamically are added to a near-saturated solution of coformer
Strategies for Low Drug Solubility 363

(i.e., the green arrow in Fig. 19). As drug is dissolved in ½AB ½AB
K 11 5 5 ð27Þ
the coformer solution, the concentration of drug in ½A½B K sp
solution passes the phase solubility limit of the
cocrystal (where the green arrow crosses the red line, The concentration of complex in solution is given by:
marked Y in Fig. 19), resulting in precipitation of
cocrystal. This leads to a reduction in coformer (and ½AB5K 11 K sp ð28Þ
drug) concentrations in solution shifting the position
on the phase diagram to the left (since coformer and the total concentration of A in solution (AT) is
precipitates) and down (since drug also precipitates a function of the solution concentration resulting from
with the cocrystal). However, in the presence of excess the solubility of the cocrystal (determined by Ksp) plus
the additional quantity, reflecting the presence of the
drug, the decrease in drug concentration is only
association complex AB (determined by K11). For a 1:1
transient as more drug immediately dissolves and
cocrystal complex in the solid state and in solution
stimulates further cocrystal precipitation. The process,
(Nehm et al., 2006):
therefore, continues, moving the composition in solu-
tion to the left of the phase diagram as drug con- K sp
centrations in solution are maintained by drug ½AT 5 1 K 11 K sp ð29Þ
½BT
dissolution but where coformer concentrations are
reduced as a result of precipitation of the cocrystal
(i.e., following the green arrow). Ultimately, the From eq. 29, a graph of [A]T against the inverse of
concentration of coformer in solution is reduced such [B]T (total coformer in solution) reveals Ksp from the
that it is no longer in excess and therefore no longer slope and K11Ksp from the intercept.
limiting to cocrystal solubility. This condition is 3. Solubility Advantages of Cocrystals. Depending on
denoted by the transition point X, which reflects the the choice of coformer, the solubility ratio of cocrystal to
solubility limit of both drug and cocrystal in solution. drug crystal may vary considerably, from less than 1 to
Alternatively, the same transition point X can be values in excess of 100. Therefore, the potential solubility
determined by adding excess cocrystal to a saturated enhancement provided by cocrystals will likely exceed
solution of drug (i.e., the blue arrow in Fig. 19). In this that of most crystalline polymorphs, and solubility
approach, dissolution of cocrystal results in increases enhancement similar to or above that of amorphous
in drug concentrations greater than the saturated drug is possible (Remenar et al., 2007; Good and
solubility, prompting drug crystallization from the Rodriguez-Hornedo, 2009). Cocrystal solubility ratios
supersaturated solution. This process again continues are also proportional to the solubility of the coformer
(Fig. 20), and data obtained for a series of cocrystals of
until the solubility of the cocrystal is reached, at which
point additional increases in drug concentration in
solution are not possible on addition of excess cocrys-
tal, and the transition point is reached (i.e., point X).
At the transition point, the existence of a mixed solid
phase (crystallized drug and cocrystal) can be verified
by x-ray powder diffraction (XRPD) and drug-coformer
concentrations in solution measured by HPLC. From
these concentrations, Ksp is calculated using eq. 26;
subsequently, the equilibrium solubility of the cocrys-
tal is calculated from Ksp. For the dissolution of a 1:1
cocrystal complex, this occurs when drug and coformer
concentrations in solution are equal and the equilib-
rium solubility of the cocrystal is equal to the square
root of Ksp.
Interestingly, solubility values predicted by Ksp, are
often lower than those obtained experimentally (Rodrí-
guez-Hornedo et al., 2006). This can be attributed to
the formation of drug-coformer complexes in solution,
which in the case of poorly water-soluble drugs (which
are usually less soluble than the coformer) leads to
increased cocrystal solubility (Nehm et al., 2006).
Where K11 is the binding constant used to describe Fig. 20. The relationship between cocrystal solubility and the solubility
of the coformer. Drugs used were caffeine, theophylline, and carbamaze-
the solution complex (AB) between drug (A) and its pine. Open symbols are solubilities determined in water, closed symbols
coformer (B) in a 1:1 ratio: in organic solvents. Adapted from Good and Rodriguez-Hornedo (2009).
364 Williams et al.

carbamazepine, theophylline, and caffeine suggest The solubility advantages provided by isolation of
that coformer solubility values must be at least 10-fold a cocrystal have been shown to correlate broadly with
higher than drug solubility to generate cocrystals with differences in melting point of the cocrystal and the
improved solubility greater than that of crystalline drug drug for a series of carbamazepine, theophylline, and
(Good and Rodriguez-Hornedo, 2009). caffeine cocrystals (Good and Rodriguez-Hornedo,
Given that cocrystal solubility is dependent on co- 2009). This trend was also replicated for 10 different
former solubility, the use of highly soluble coformers is cocrystals of AMG 517 (Stanton and Bak, 2008),
common. For ionizable coformers, this results in sig- providing evidence of the role of decreased crystal
nificant differences in cocrystal solubility with changes lattice energy in the solubility improvement. However,
in pH even for cocrystals with nonionizable drugs. For this relationship was not apparent in later investiga-
example, Fig. 21 illustrates theoretical pH solubility tions of cocrystals of AMG 517 using a more diverse
profiles for a cocrystal containing a nonionizable drug range of coformers (Stanton et al., 2009), and closer
and either a weakly acidic (Fig. 21A) or a zwitterionic scrutiny of the data sets for carbamazepine and theoph-
coformer (Fig. 21B). The corresponding effects for an ylline revealed that the decrease in melting point of the
ionizable drug in the presence of both coformers are cocrystal was not sufficient to explain the magnitude of
also illustrated (Fig. 21, C and D) (Bethune et al., the increase in solubility (Good and Rodriguez-Hornedo,
2009). Note that ionizable drugs may form cocrystals 2009). These data are consistent with the view that
(rather than salts) with ionizable coformers where the enhanced solute-solvent interactions (such as the forma-
difference in pKa between drug and conformer is tion of drug-coformer complexes in solution) and the use
insufficient to allow salt formation. of coformers with increasing aqueous solubility may be

Fig. 21. Theoretical pH solubility profiles of various ionizable and nonionizable compounds and their cocrystals. (A) carbamazepine-succinic acid, (B)
carbamazepine-4-aminobenzoic acid hydrate, (C) itraconazole-L-tartaric acid, and (D) gabapentin-3-hydroxybenzoic acid. The plots show that even for
nonionizable drugs [i.e., (A) and (B)], pH can affect the solubility of the cocrystal via effects on the solubility of the coformer. Adapted from Bethune
et al. (2009).
Strategies for Low Drug Solubility 365

more significant determinants of increased cocrystal and the difficulties associated with distinguishing
solubility than changes to solid-state properties (Good between the interactional hierarchies that commonly
and Rodriguez-Hornedo, 2009; Schultheiss and Newman, exist when the drug contains multiple functional
2009). groups with the capacity of forming multiple supramo-
Because the dissolution of a cocrystal with higher lecular synthons with a coformer (Haynes et al., 2004;
solubility than drug alone leads to the creation of Aakeroy and Salmon, 2005). Therefore, both computa-
a solution that is supersaturated with respect to drug tional and empirical studies are required to support
solubility, the biopharmaceutical advantages of cocrys- cocrystal design (Shan and Zaworotko, 2008; Arora and
tals are based on increases in kinetic or apparent Zaworotko, 2009), and it is common to screen for
solubility. The stability of this supersaturated solution potential cocrystal hits using a range of coformers.
is therefore of paramount importance. The generation These may be chosen preferentially from coformers
(and stabilization) of supersaturated drug solutions is sharing the same functional groups (Remenar et al.,
an important aspect of the performance of many 2003b; Childs et al., 2004; Childs and Hardcastle, 2007;
contemporary formulation approaches, including solid Stanton et al., 2011) but might also include, for
dispersions and lipid-based formulations, and is example, a mixture of acids and amides to explore
addressed elsewhere in this review (see Sections X whether assemblies of homosynthons or heterosyn-
and XI, respectively). Regardless of the manner in thons are preferred (Fleischman et al., 2003; Cheney
which supersaturation is generated (often described as et al., 2010).
the “spring” required to generate supersaturation), Cocrystals can be prepared from solutions containing
a means of reducing the rate of recrystallization of the stoichiometric amounts of drug and cocrystal former
stable crystal form (i.e., a “parachute”) is important. In followed by removal of the solvent through evaporation
the case of cocrystals, this may be achieved by judicious or by cooling the mixture to induce precipitation of the
choice of cocrystal designs that do not necessarily solid phase (Etter and Reutzel, 1991; Fleischman et al.,
promote the highest kinetic solubility (since this also 2003; Walsh et al., 2003; Childs et al., 2004). Other
generates the greatest driving force for recrystalliza- common preparation methods include ball mill cogrind-
tion) but rather provide the slowest rate of trans- ing, antisolvent addition, melt cooling, sublimation,
formation to the less-soluble form. Alternatively, ultrasonication, slurry conversion, and supercritical
formulation additives may be used in an attempt to technologies (Morissette et al., 2004; Arora and Zawor-
stabilize the supersaturated solution; to this end, otko, 2009; Padrela et al., 2009). More recently, the
surfactants and polymers have recently been explored solvent-drop method has been increasingly used, in
as a means of stabilizing supersaturated cocrystal which small amounts of solvent are added during
solutions (Remenar et al., 2007; Huang and Rodriguez- cogrinding of the drug and coformer (Trask et al.,
Hornedo, 2010; Good et al., 2011). 2004). Solvent methods are limited by virtue of the often
4. Design and Preparation of Cocrystals. Tra- wide differences in solubility of the drug and coformer
ditionally, pharmaceutical cocrystals have been formed (Morissette et al., 2004; Qiao et al., 2011), but they offer
via empirical attempts to cocrystallize a drug of interest the advantage of more rapid rates of crystallization and
with a wide range of pharmaceutically approved or the absence of physical and thermal stresses. In general,
acceptable cocrystallizing agents or coformers. The however, these techniques do not assure isolation of the
advent of in silico crystal engineering methods now cocrystal and may result in crystallization of the single
provides an opportunity for more rational cocrystal crystal phases of drug or conformer alone. Phase
design. Central to this design step is the identification of diagrams are beneficial in this regard as they reveal
intermolecular interactions, or synthons, that form the solvent:drug:coformer ratios at which the most
between the drug and a potential cocrystal former. stable phase is the cocrystal (Chiarella et al., 2007;
These interactions ultimately lead to the formation of Ainouz et al., 2009). This is particularly important when
supramolecular assemblies and a crystal structure when the drug and the coformer show wide differences in
extended into three-dimensional space. Synthon selection solubility in the solvent (Blagden et al., 2007; Chiarella
during cocrystal design has been aided significantly by et al., 2007). An understanding of the phase solubility
large number of cocrystal structures listed in the Cam- behavior of drug, coformer, and solvent underpins the
bridge Structural Database (Allen, 2002; Shan and reaction crystallization method (RCM) of cocrystal
Zaworotko, 2008). formation where excess drug is added to saturated or
Although the rational design of the supramolecular near-saturated solutions of coformer. The RCM is
synthon provides a starting point for cocrystal assem- described in more detail in Section V.B.2 and in Fig.
bly, it is not always possible to translate this into 19. The methods used to prepare cocrystals can strongly
reality (Aakeroy and Salmon, 2005). Indeed, factors influence the possibility of success, and a large number
other than the theoretical molecular compatibility of potential coformers can be investigated. Conse-
between a drug and a cocrystal former may dominate, quently, several studies have presented high-
including the intrinsic unpredictability of crystallization throughput screening approaches to cocrystal synthesis
366 Williams et al.

and physicochemical characterization (Remenar et al., these articles originate from the last 5 years. Table 13
2003a; Morissette et al., 2004; Seefeldt et al., 2007; provides a summary of most of the recent examples,
Zhang et al., 2007; Childs et al., 2008; Lu et al., 2008; and the following section provides examples of a more
Kojima et al., 2010). limited subset where in vitro evaluation of solubility
As with all solid-state screening methods, the and dissolution has been progressed to in vivo
isolated solid phase must be characterized, and confirmation of a bioavailability advantage.
common spectroscopic techniques used to verify the Several studies have examined the ability of AMG
presence of a cocrystal include XRPD, NIR, FTIR, and 517 (a poorly soluble vanilloid receptor-1 antagonist for
Raman Spectroscopy. The presence of a single melting the treatment of acute and chronic pain) to form
endotherm by DSC has been widely used to verify the pharmaceutical cocrystals (Bak et al., 2008; Stanton
presence of a cocrystal. Single-crystal X-ray diffrac- and Bak, 2008; Stanton et al., 2010; Stanton et al.,
tometry is also increasingly being used to determine 2011). Interestingly, the potential to form cocrystals
structural details of the cocrystal (McNamara et al., with AMG 517 [N-(4-[6-(4-trifluoromethyl-phenyl)-
2006; Remenar et al., 2007), and NMR and HPLC are pyrimidin-4-yloxy]-benzothiazol-2-yl)-acetamide] was
being used to verify the stoichiometric ratio of the first discovered serendipitously when a cocrystal with
complex (Bak et al., 2008). An additional characteriza- sorbic acid was identified in an oral suspension contain-
tion step is required where drug and coformer have the ing sorbic acid as a preservative. In these initial studies,
potential to undergo an acid-base reaction to form the in situ formation of the cocrystal in the original
a salt rather than a cocrystal. By definition, compo- suspension led to good in vivo exposure (in rats), albeit at
nents of cocrystals are neutral. However, many drug relatively low doses. Studies at much higher doses (up to
compounds contain potential sites of ionization that 500 mg/kg) subsequently confirmed the utility of the
may cooperate with an oppositively charged group on cocrystal approach and showed good exposure from
the coformer. To identify whether a salt or cocrystal suspensions containing stoichiometric ratios of the drug
results, the position of the proton of interest must be and the sorbic acid coformer (Bak et al., 2008). AMG 517
characterized, most commonly via single crystal X-ray cocrystals were later prepared using a range of carboxylic
diffractometry, solid-state NMR, or X-ray photoelec- acids and amide-containing coformers (Stanton et al.,
tron spectroscopy (Aakeroy et al., 2007; Stevens et al., 2010; Stanton et al., 2011), and each cocrystal showed
2010; Mohamed et al., 2011). superior rates of dissolution in comparison with AMG
5. Recent Examples of Pharmaceutical Cocrystals 517 and increased bioavailability after oral administra-
for Improving Solubility and Bioavailability. Cocrystals tion to rats. In these studies, cocrystals were dosed in
have been isolated for a range of poorly water-soluble suspensions supplemented with 1% PVP to reduce the
drugs in an attempt to increase kinetic solubility and rate of recrystallization of drug (rather than cocrystal)
in vivo exposure, although the relatively recent ap- since this process was shown, in vitro, to be rapid (,1 h
preciation of the use of cocrystals as a means of for most of the investigated cocrystals) (Stanton et al.,
overcoming low drug solubility dictates that most of 2010).

TABLE 13
Cocrystal references describing effects on in vitro dissolution rate and in vivo bioavailability

Drug Coformer(s) Effect of Cocrystallization on Drug Dissolution


and/or In Vivo Exposure

Exemestane (Shiraki et al., 2008) Maleic acid A faster intrinsic dissolution rate compared with drug alone, although
enhancement in dissolution rate was limited by a rapid rate of
transformation of the cocrystal to the drug in solution.
Megestrol acetate (Shiraki et al., 2008) Saccharin A 3.8-fold increase in dissolution rate generating supersaturation but with
rapid (within 20 min) transformation and recrystallization of drug alone.
After 4 h, concentrations of dissolved drug were 2-fold greater than the
equilibrium solubility of the drug alone.
Indomethacin (Jung et al., 2010) Saccharin A 1.7- and 3.6-fold increase in intrinsic dissolution at pH 1.2 and 7.4,
respectively.
Approximately a 2-fold increase in AUC in beagle dogs and a .6-fold
increase in Cmax.
Meloxicam (Cheney et al., 2011) Aspirin A .40-fold increase in kinetic solubility.
A . 4-fold increase in AUC in rats.
Nitrofurantoin (Cherukuvada et al., p-Aminobenzoic The PABA cocrystal showed faster (.1.4-fold) intrinsic dissolution rate
2011) acid (PABA) compared with drug alone.
Urea The dissolution rate of an arginine salt was faster than the urea cocrystals
but was more susceptible to hydrate formation.
Quercetin (Smith et al., 2011) Isonicotinamide Up to 14-fold increase in kinetic solubility, with evidence of rapid (,1 h)
recrystallization.
Caffeine AUC in rats increased nearly 10-fold, with reduction in Tmax.
Theophylline Authors were unable to directly correlate the in vitro dissolution rate of
Theobromine cocrystals with in vivo performance.
Tmax, time after administration that a drug is at its maximal concentration.
Strategies for Low Drug Solubility 367

Cocrystal formulations have been described for the bioavailability was also observed in fasted beagle dogs
poorly water-soluble sodium channel blocker, CFPC compared with the drug crystal when the cocrystal was
[2-[4-(4-chloro-2-fluorophenoxy)-phenyl]pyrimidine-4- dosed at 5 or 50 mg/kg (McNamara et al., 2006) (Fig. 22B).
carboxamide]. CFPC has a low pKa, which precludes Cocrystal strategies have also been used to improve
facile salt formation with most acidic counterions, and the in vivo pharmacokinetic profile (increased expo-
a low glass transition temperature (Tg 5 43°C), sure, faster onset of action) of L-883555 (a novel
rendering it highly sensitive to recrystallization from phosphodiesterase-IV inhibitor complexed with L-
amorphous formulations (McNamara et al., 2006). tartaric acid) (Variankaval et al., 2006), indometha-
cin (with saccharin) (Jung et al., 2010), and melox-
Cocrystal approaches to solubility enhancement were
icam (with aspirin) (Cheney et al., 2011). In the last
therefore attempted, and a glutaric acid containing
example, an unconventional approach to cocrystal
cocrystal was selected for in vitro and in vivo analysis.
design was used where meloxicam-aspirin cocrystals
In comparison with the free base, the glutaric acid were successfully formed after only consultation of the
cocrystal showed an 18-fold increase in intrinsic disso- Cambridge Structural Database and a priori knowledge
lution rate (Fig. 22A). Consistent with the increase in in of a pharmacological interaction between these two
vitro dissolution rate, a 2.5- to 3.3-fold increase in compounds [coadministration of commercial meloxicam
(Mobic; Boehringer Ingelheim, Ridgefield, CT) and
aspirin had previously led to an unexplained increase
in AUC] (Cheney et al., 2011).

C. Summary
Polymorphism in pharmaceutical solids is common.
Decreases in intermolecular attractive forces in high-
energy polymorphs and the amorphous form (compared
with the stable crystal form) dictate that these
materials exhibit higher apparent solubilities and have
the potential to improve dissolution rate, solubilization,
and absorption for poorly water-soluble drugs. However,
the use of high-energy polymorphs (or more commonly
the amorphous form) is complicated by thermodynamic
instability in the solid state. Additional approaches
must therefore be taken to stabilize formulations to
provide acceptable shelf-life. This is commonly achieved
by combination with excipients such as polymers to
form solid dispersion formulations (these are described
in more detail in Section X).
Crystal-engineering strategies provide an additional
approach for the manipulation of crystal habit and an
additional opportunity to reduce the strength of the
crystal lattice and to promote solubility and absorption
for poorly water-soluble drugs. Cocrystals constitute
a molecular complex between a drug and a water-soluble
cocrystal former and are in some respects structurally
similar to salts. In contrast to salts, however, the complex
between drug and cocrystal former does not involve
proton exchange; therefore, cocrystals can be formed by
nonionizable drug molecules.

VI. Cosolvents
A. Rationale for the Use of Cosolvents for the
Solubilization of Poorly Water-Soluble Drugs
Cosolvents are water-miscible organic solvents and
Fig. 22. Comparison of CFPC and its glutaric acid cocrystal during (A)
intrinsic dissolution testing in pure water (37°C with USP paddles at 100 are widely used to increase the solubility of poorly
rpm) and in (B) beagle dog oral pharmacokinetic testing (dosed at 50 mg/ water-soluble substances (Yalkowsky, 1999; Rubino,
kg). Values are expressed as means [n 5 3 in (A) and n 5 6 in (B)] 6 SD.
CFPC is 2-[4-(4-chloro-2-fluorophenoxy)-phenyl]pyrimidine-4-carboxa- 2002). Cosolvents are less polar than water and,
mide. Adapted from McNamara et al. (2006). therefore, enhance aqueous drug solubility by lowering
368 Williams et al.

the polarity of the bulk solvent to a level that more closely solubility in the formulation. These formulations are
reflects the polarity of the nonpolar solute (e.g., drug). described in more detail in Section XI.
Cosolvent formulations are used routinely for the The focus of the present section is on the utility of
parenteral delivery of drugs during preclinical assess- cosolvents in enhancing drug solubility for parenteral
ment and are widely used in marketed parenteral delivery where the most commonly used cosolvents
products, including diazepam (Valium; Hoffmann-La include ethanol, propylene glycol, and low-molecular-
Roche, Inc., Nutley, NJ), digoxin (Lanoxin; Glaxo- weight polyethylene glycol (PEG). Here we describe the
SmithKline, Research Triangle Park, NC), melphalan fundamental concepts that underpin the solubilization
hydrochloride (Alkeran; GlaxoSmithKline), and lora- potential of cosolvents, the potential of the log-
zepam (Ativan; Wyeth-Ayerst, Radnor, PA). Their solubility model to predict drug solubility in cosolvent
popularity can be attributed to the relative ease of systems, the key formulation/nonformulation factors
use in formulation design and the potential to increase that can affect solubilization, and, finally, the potential
dramatically the level of solubilized drug. pharmacological effects of cosolvents.
Drug candidates most well suited to cosolvency
include those that lack ionizable functionalities (and B. Commonly Used Organic Cosolvents in Parenteral
therefore are not suitable for pH adjustment strate- Drug Delivery
gies) and those with moderate log P (between 1 and 3) Cosolvents are used in 10 to 15% of FDA-approved
that show relatively low affinity for solubilization by parenteral products (Nema et al., 1997), and the most
surfactants or lipids (Kipp, 2007). In addition, cosol- widely used cosolvents are propylene glycol, ethanol,
vents are commonly used in combination with other glycerol, and PEG 300/400. Examples of the cosolvent
solubilization strategies. For example, the use of a cosol- concentrations currently used in marketed formula-
vent with a surfactant can allow high concentrations of tions are shown in Table 14, where it is apparent that
solubilized drug to be obtained (via discrete mechanisms) the concentrations of cosolvent used are often high and
at lower concentrations of each excipient. Surfactants are many formulations contain .50% (v/v) cosolvent [e.g.,
also widely used in conjunction with cosolvents to reduce Prograf (Astellas Pharma US, Inc, Deerfield, IL),
the risk of drug precipitation as cosolvent-based for- melphalan (Alkeran; GlaxoSmithKline, Research Tri-
mulations are diluted in vitro or in situ. angle Park, NC), and lorazepam (Ativan; Wyeth-
Cosolvents are also widely used in formulations Ayerst, Radnor, PA). In this case, formulations for i.v.
designed for oral delivery. In this case, however, administration are diluted before use to minimize
cosolvents are commonly a minor component of the irritation at the injection site (see Section VI.F).
formulation, and other excipients provide the major However, dilution of the formulation will significantly
solubilization enhancement. Lipid-based formulations reduce the solubilization capacity of the cosolvent,
are typical of systems of this type and may comprise thereby introducing the risk of drug precipitation.
lipids, surfactants, and cosolvents. Under these cir- Alongside the possibility of incompatibility or toxicity,
cumstances, cosolvents may be included to reduce the potential for drug precipitation either on dilution
viscosity, to facilitate dispersion, and to enhance drug before administration or on injection is a key criterion

TABLE 14
Commonly used pharmaceutical cosolvents in parenteral drug delivery: recommended concentrations in preclinical testing and maximum
concentrations observed in clinical use

Cosolvent Recommended Range Frequency of Use in Concentrations Used in Some Final Cosolvent Concentration
for Preclinical Usea Commercial Formulationsb Commercial Formulationsb Administered (%)b

Glycerol 17 70% in Multitest CMI Not diluted before s.c. use.


Propylene glycol 30–60% 32 80% in Ativan #40% for i.v. bolus dose
40% in Dilantin Direct i.v. injection with a maximum
rate of administration of 1 ml/min.
40% in Valium Direct i.v. injection with a maximum
rate of administration of 1 ml/min.
40% in Lanoxin #25% before i.v. infusion
68% in Luminal #68% before i.v. bolus
60% in Alkeran #1% before i.v. infusion
Ethanol 10% 34 100% in alprostadil #4% for i.v. infusion
injection USP
49% in Taxol #10% before i.v. infusion
80% in Prograf #0.32% before i.v. infusion
42% in Vumon #0.84% before i.v. infusion
Polyethylene glycol 300 4 50% in methacarbamol Direct i.v. injection with a maximum
rate of administration of 3 mL/min.
Polyethylene glycol 400 40–100% 4 67% in Busulfex #5.6% before i.v. infusion
DMSO 10–20%
DMA 10 –30% 2 33% in Busulfex #2.75% before i.v. infusion
a
% w/v, on the basis of single dosing 10 ml/kg to a mouse or rat. Li and Zhao (2007). See text for additional details.
b
Values obtained from Nema and Brendel (2011), Strickley (2004), and Bouqet and Wagner (2012).
Strategies for Low Drug Solubility 369

in the design of parenteral cosolvent formulations. As also contains PEG 400 but with glycerin, citric acid, and
lower formulation volumes are typically used during a small amount of water; amprenavir capsules and oral
i.m. administration, higher cosolvent concentrations solution (Agenerase; GlaxoSmithKline), which contain
(i.e., lower dilutions) may be used if the formulation is PEG 400, propylene glycol, TPGS, sodium chloride,
intended for this route (Gad et al., 2006). Table 14 sodium citrate, and citric acid.
provides some guidance on the recommended concen-
trations of cosolvents for preclinical use (Li and Zhao, C. Solubilization by Cosolvents
2007). Although formulations typically do not contain
1. Cosolvent Effects on Solubility Parameters.
ethanol at greater than 10% v/v, because of potential
Cosolvents increase the aqueous solubility of poorly
central nervous system and respiratory depressive
water-soluble drugs by disrupting the intermolecular
effects, there are many exceptions. For example,
hydrogen bonding networks that are present in
ethanol concentrations between 15% and 20% v/v have
aqueous systems. This leads to a decrease in the
been administered intravenously to canines (Chang
polarity of the solvent and the creation of an environ-
et al., 2007; Gu et al., 2009); in rats, ethanol
ment with physicochemical properties that are more
concentrations up 80% have been used (Hanafy et al.,
similar to that of the solute (e.g., drug). Simplistically,
2007). Recommended concentrations of propylene
this results in an increase in drug solubility according
glycol and PEG 400 that may be administered in-
to the general principle of “like dissolves like.”
travenously are usually much higher than ethanol
More accurately, the effect of cosolvents on drug
(Table 14). Propylene glycol has been administered
solubility may be assessed using the solubility param-
intravenously to beagles at concentrations between 40
eter approach. The role of solubility parameters in
and 60% (Joshi et al., 2004; Merritt et al., 2007), and
ideal and nonideal solubility was discussed in Section
formulations containing up to 100% propylene glycol
I.B.1 of this review. In brief, differences between the
(PG) have been administered to rats (Gershkovich and
solubility parameters of solute (drug) and solvent
Hoffman, 2007; Usach and Peris, 2011). A recent study
provide an indication of the differences in intermolec-
reported incidences of reversible renal toxicity and
ular forces in the solute (i.e., solute-solute interactions)
increased water consumption/dry mouth in dogs that
and solvent (i.e., solvent-solvent interactions) and in
were intravenously dosed with up to 8.75 g/kg PEG 400
turn provide an indication of how close the properties
per day over 30 consecutive days, confirming that this
of the solution are to an ideal solution (eq. 8).
cosolvent is generally well tolerated (Li et al., 2011a).
In an ideal solution, intermolecular forces between
Dimethyl sulfoxide (DMSO) is not used in commer-
solute-solute and solvent-solvent molecules are equal,
cial parenteral formulations because of the potential
and under these conditions, solubility is essentially
for membrane penetration and cell lysis and other
maximal (and solvent independent). However, the
potential hemodynamic disturbances (e.g., clotting,
large differences between intermolecular bonds in
changes to arterial blood pressure) (Mottu et al.,
water and most poorly water-soluble solutes dictate
2000). Nonetheless, DMSO is widely used in preclinical
that almost all aqueous solutions are nonideal. Devia-
formulations (usually up to 20% v/v) (Vennerstrom
tions from ideality result in a decrease in solubility,
et al., 2004; Liddle et al., 2008; Caliph et al., 2009;
and the magnitude of the deviation from ideality is
Kapetanovic et al., 2010; Usach and Peris, 2011), and
captured by the activity coefficient (g) (eq. 7), which in
a DMSO containing busulfan intravenous formulation
turn is a function of the difference in the solubility
(containing up to 2% v/v DMSO, equivalent to #0.5 ml)
parameters of solute and solvent (eq. 8).
has been tested in canine models without showing any
The solubility parameter (d) may be determined
signs of adverse effects (Ehninger et al., 1995) and in
experimentally by measuring the energy required to
human subjects receiving a single i.v. dose who were
vaporize (Ev) a solute or solvent of molar volume Vm
also free from pain at the injection site (Schuler et al.,
where
1998). Cosolvent combinations are widely used in pre-
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
clinical i.v. formulations. Ethanol has been frequently
used in combination with propylene glycol (Kamath
d5 DEv
= Vm
ð30Þ
et al., 2005; Zhang et al., 2006; Otaegui et al., 2009) or
PEG 400 (Stella et al., 1995; Chan et al., 1998; Choo More commonly, however, solubility parameters are
et al., 2011); DMSO has been used with propylene estimated using the group contribution approach, as the
glycol combinations (Chen et al., 2007; Hu et al., 2009; total cohesive forces of a molecule reflect the cohesive
Usach and Peris, 2011). properties of the individual nonpolar and polar frag-
Examples of oral products that contain significant ments (Fedors, 1974). The contributions to total cohesive
amounts of cosolvent include liquid-filled bexarotene force (dt) of dispersive forces (dd), the polar permanent
capsules (Targretin; Eisai Inc., Woodcliff Lake, NJ), dipole (dp), and hydrogen bonding forces (dh) are
which use PEG 400 as the cosolvent; etoposide capsules provided by the Hansen partial solubility parameters
(VePesid; Bristol-Myers Squibb, Stamford, CT), which (eq. 31):
370 Williams et al.

TABLE 15
Examples of cosolvents widely used in parenteral drug delivery and corresponding descriptors of their relative polarity
Cosolvent Solubility Parametera (d) Dielectric Constanta («) Interfacial Tensiona Log Kowa

cal/cm3 dynes/cm
Water 23.4 81.0 72.0 24.00
Glycerol 16.5 42.5 64.9 22.60
Ethylene glycol 14.6 37.7 48.8 21.93
Propylene glycol 12.6 37.7 37.1 21.40
Ethanol 12.7 24.3 22.2 20.31
DMSO 12.0 46.7 38.0 21.09
PEG 400 11.3 13.6 46.0a, 11.7b 21.30
DMA 10.8 37.8 35.7 20.66
Dioxane 9.9 2.2 33.6 20.42
Kow, the oil:water partition coefficient.
a
Values obtained from Yalkowsky (1999).
b
Values obtained from Rubino (2002).

qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
dt 5 d2d 1 d2p 1 d2h ð31Þ ratio for combinations of specific solvents (see the
following section below).
Experimentally calculated solubility parameters for Other commonly used descriptors of cosolvent polar-
most cosolvents are readily available in the literature, ity include interfacial tension and oil-water partition
and values for some of the most commonly used coefficient. Values of these parameters for commonly
cosolvents are shown in Table 15. As a general rule, used cosolvents are also presented in Table 15.
the solubility parameter decreases with decreasing 2. Predicting Solubility in Cosolvent Systems: The
cosolvent polarity as the capacity to self-associate via Log-Linear Solubility Model. A number of models have
hydrogen bonding is reduced. In contrast, water has been described that aim to predict drug solubility in
the highest solubility parameter. The solubility of cosolvent combinations, and for further information,
nonpolar solutes (the solubility parameters for which the interested reader is directed to a recent review by
are low relative to that of water) is therefore higher in Jouyban (Jouyban 2008). Of these, the log-linear
cosolvents or in mixed cosolvent-water solutions with solubility model, proposed by Yalkowsky, is perhaps
lower solubility parameters. the most widely used (Li et al., 1999b,d; Yalkowsky,
The dielectric constant («), defined by the ability of 1999; Millard et al., 2002; Kawakami et al., 2004;
a material to concentrate electric lines of flux, also Kawakami et al., 2006; Tomsic et al., 2006; Dong et al.,
provides a measure of cosolvent polarity. In the 2008). The log-linear model suggests that the molar
context of cosolvent systems, dielectric properties solubility of a solute in a mixed solvent system is a
provide a measure of the ability of the solution to function of drug solubilities in the individual compo-
allow electrolyte dissociation into respective ions. nents (Yalkowsky, 1999; Millard et al., 2002). Equation
Water, for example, has a high dielectric constant, 32 describes the relationship between the solubility of
readily allows ionic dissociation, and therefore is a solute in a binary water-cosolvent system (Smix),
a good solvent for polar and ionizable compounds. In where Sw is the molar solubility of the solute in water, fc
contrast, the high dielectric constant and polarity of is the volume fraction of the cosolvent, and s is the
water result in lower solubility for nonpolar mole- solubilization power of the cosolvent. Equation 33 may
cules. Dioxane, by comparison, is miscible with water be further used to estimate drug solubility in a system
but has a very low dielectric constant (« 5 2.25, containing multiple cosolvents, although it assumes
Table 15). Blending dioxane with water in differing that the different cosolvents exhibit no specific nonideal
ratios, therefore, generates a series of standard interactions (i.e., that the solubilization powers of the
different solvents are simply additive) (Millard et al.,
solutions with known dielectric constant. Evaluation
2002):
of the solubility of a solute in these dioxane-water
mixtures allows the dielectric constant required for logSmix 5logSw 1 sf c ð32Þ
maximum solubility to be identified [the dielectric
requirement (DR)]. Since maximum solubility will logSmix 5logSw 1 +sf c ð33Þ
occur at the same (or within a narrow range) DR value
in other cosolvent systems, the DR value can be used Knowledge of the intrinsic solubility of the solute in
to calculate the proportion of cosolvent(s) required to water and the solubilization power of the cosolvent (s)
achieve maximum solubility in a defined solvent- can therefore be used to predict drug solubility at
cosolvent pair (Paruta et al., 1964). Importantly, a particular cosolvent concentration. The value for s
however, the solubility value at this maximum will may also be calculated using eq. 34, where s and t are
vary depending on the nature of the cosolvent pairs constants for a particular solvent (but are solute
used; hence, the DR is able to provide only the ideal independent) and have been determined for a variety
Strategies for Low Drug Solubility 371

of commonly used cosolvents using experimental polarity of the cosolvent and supports the general
solubility data for different polar and nonpolar com- suggestion that less polar solvents are typically more
pounds (Millard et al., 2002). Kow is the oil-water effective solvents for nonpolar solutes. However, for
partition coefficient for the solute: many of the other examples in Table 16, ethanol
provides improved solubilization compared with PEG
s5slogK ow 1 t ð34Þ 400 and DMSO, despite a higher dielectric constant
and solubility parameter. It is also apparent that the
An example of the applicability of the log solubility rank order of solubility properties changes with pH
model is provided in Fig. 23, which plots the solubility (Jain et al., 2001). Measures of cosolvent polarity alone
of a series of alkyl p-aminobenzoates as a function of are therefore not always predictive of solvent capacity,
propylene glycol concentration in water (Yalkowsky and other factors—including potential hydrogen bond-
et al., 1972). For each solute, excellent agreement is ing between the drug and solute—should also be
evident between the experimentally determined solu- considered.
bility (symbols) and predicted solubility (solid lines) A limitation to the use of the log-linear solubility
according to eq. 32. It is also apparent that the effect of model is the lack of applicability for semipolar solutes
cosolvent concentration (i.e., solubilizing power) be- (i.e., those solutes that are more polar than cosolvent
comes increasingly more pronounced with increasing but less polar than water). These compounds show a
alkyl chain length (and therefore hydrophobicity) of the solubility maximum at cosolvent mixtures between
solute, as suggested by eq. 34. Indeed, near-linear 0 and 100% cosolvent. This is in contrast to the more
relationships have been identified between the solubi- common scenario of polar or nonpolar solutes (where
lization power (s) of many commonly used cosolvents the log-linear model appears to have broader applica-
(ethanol, propylene glycol, PEG 400, glycerol) and the bility), where solubility is greatest in 100% water or
hydrophobicity of the solute (as described by log Kow) 100% cosolvent, respectively (Morris et al., 1988;
(Millard et al., 2002). Yalkowsky, 1999; Millard et al., 2002; Rubino, 2002).
Since s and t are constants for a particular cosolvent, Experimentally determined log solubility versus cosol-
eq. 34 also predicts that the rank order of solubilization vent fraction plots also often show deviation from the
power for a range of cosolvents should be consistent model at high cosolvent concentrations, where mea-
across different solutes. However, this is not always sured solubility is higher or lower than the solubility
the case, and the literature provides many examples predicted by the linear log model. Negative deviations
where rank-order solubility for different cosolvent (i.e., solubility lower than predicted) are more preva-
systems is highly compound specific (Table 16). For lent at low cosolvent concentrations (e.g., ,50%) and
example, solubilization of carbendazim at pH 7 (where may be due to the ability of water to retain its ordered
the drug is predominantly un-ionized) follows the structure and the formation of some water-cosolvent
rank-order PEG 400 . ethanol . propylene glycol . intermolecular hydrogen bonds, in turn decreasing the
glycerol. In this case, solubility reflects the changing potential number of solvent-solute interactions

TABLE 16
Rank order in drug solubilization obtained in various water/cosolvent
binary mixtures

Drug Rank Order in Solubilization Power


(s)a

Antalarmin (Sanghvi et al., Ethanol . PG . PEG 400


2009)
Phenytoin (Kawakami et al., DMA . ethanol 5 PEG 400 .
2004, 2006). DMSO
NSC-639829, pH 7 Ethanol . DMSO . PEG 400 .
(un-ionized) PG
NSC-639829, pH 2 Ethanol . PEG 400 . PG .
DMSO
NSC-639829, pH 1 (ionized) PG . ethanol . PEG 400 .
(Jain et al., 2001) DMSO
Carbendazim (Ni et al., PEG 400 . ethanol . PG .
2002) glycerol
N-Epoxymethyl-1,8- Ethanol . PEG 400 . PG .
naphthalimide glycerol
(Dong et al., 2008)
PG-300995 (Ran et al., Ethanol . PEG 400 . PG .
Fig. 23. Log-linear solubility relationship for a series of alkyl p-amino- 2005) glycerol
benzoates in propylene glycol-water mixtures. A decreasing slope reflects PG, propylene glycol.
decreasing log P values for the various solutes. Alkyl chain lengths on the a
Solubilization was determined over a 0–20% w/v cosolvent range, with the
drug were ethyl (C2), butyl (C4), hexyl (C6), octyl, (C8), dodecyl (C12). exception of studies by Jain et al. (2001) and Dong et al. (2008), where maximum
Adapted from Yalkowsky et al. (1972). cosolvent concentrations were 40–60% and 50%, respectively.
372 Williams et al.

(Kimura et al., 1975; Rubino and Obeng, 1991). At of poorly water-soluble drugs may be enhanced (see
higher cosolvent concentrations (e.g., .50%), there is Section VII), and cosolvent-surfactant combination
a significant breakdown of ordered water structure. strategies have been widely explored. Kawakami and
More water molecules are therefore available to coworkers, for example, showed that combining DMA
participate in solute-solvent interactions, potentially or DMSO with a nonionic surfactant, Gelucire 44/14
leading to positive deviations (i.e., solubility higher (Gattefossé Pharma, Lyon, France) led to a moderate
than predicted) in log solubility plots (Rubino and increase in the solubility of both phenytoin and
Yalkowsky, 1987; Rubino and Obeng, 1991). Alterna- indomethacin and that the solubility of phenytoin
tively, these positive deviations from predicted solu- increased with the addition of Tween 80 or sodium
bility have been attributed to changes to the solid-state dodecyl sulfate (SDS) to a variety of cosolvent formula-
drug properties, in particular, the formation of more tions (ethanol, PEG 400, glycerol or DMA) (Kawakami
soluble solvates (Flynn et al., 1979; Nicklasson and et al., 2004, 2006). In all cases, however, the additive
Brodin, 1984). effect on total solubility was small. The lack of a more
significant increase in solubility is first due to the
D. Impact of Solubilizers and Electrolytes on ability of the cosolvent to increase the aqueous
Cosolvent-Mediated Drug Solubilization solubility of the surfactant, thereby increasing the
1. Cosolvents and pH Adjustment. Buffers are CMC and decreasing micelle volume, both factors that
widely used to improve the solubilization of ionizable reduce the solubilization potential of the surfactant
compounds (see Section IV). In the presence of (see Section VIII). Second, cosolvent is expected to
a cosolvent, the total solubility of an ionizable solute partition into the micelle, leading to an increase in
is a function of the concentration of both the un-ionized polarity of the micelle core (rendering it less capable of
and ionized species and the solubilization power of the solubilizing hydrophobic drugs) and reducing the ef-
cosolvent toward the un-ionized (su) and ionized forms fective extramicellar concentration of cosolvent.
(si). Since cosolvent solubilization power is dependent However, despite the limited advantage of surfactant-
on the hydrophobicity of the solute (eq. 34), the level of cosolvent combinations on total solubilization, surfac-
cosolvency decreases with increasing drug ionization, tants have the additional (and highly significant)
potentially reducing the solubilization advantage of the benefit of reducing the potential risk of drug pre-
cosolvent. cipitation on dilution of cosolvent formulations (Li and
This effect is illustrated in Fig. 24, where increasing Zhao, 2007; Jain et al., 2010; Tonsberg et al., 2011).
concentrations of a variety of cosolvents (ethanol, This protective effect reflects the exponential decrease
propylene glycol, DMSO, and PEG 400) increase the in the solubilizing power of cosolvents with increasing
solubility of NSC-639829 (a poorly soluble weak base), dilution, whereas dilution of surfactant micelles above
but the effect (power) of each cosolvent diminishes with the CMC leads to a more linear solubility-dilution
decreasing pH (as drug ionization increases) (Jain profile. Surfactant-cosolvent combinations are there-
et al., 2001). However, the reduction in solubilization fore widely used, even though the effects on total
capacity with increased drug ionization is not sufficient solubility are only moderate (see also Section VII.D.1).
to cause a decrease in the total drug solubility, and in 3. Cosolvents and Cyclodextrins. Cyclodextrins, in-
all cases, the beneficial effect of increasing ionization cluding hydroxypropyl-b-cyclodextrin and sulfobutylether-
on solubility outweighs the reduction in solubilization b-cyclodextrin, have been used in the development of
power of the cosolvent. Thus, a solubility advantage of parenteral formulations since they can increase the
a combination strategy of cosolvent plus pH manipu- apparent solubility of a number of poorly water-soluble
lation is evident, although the beneficial effects of pH drugs by forming drug-cyclodextrin inclusion com-
adjustment are reduced at high cosolvent fractions. As plexes (see Section VIII).
such, cosolvent-pH buffering combination strategies In general, the combination of cyclodextrins with
are widely used in the development of preclinical cosolvents is uncommon, as several factors may con-
parenteral formulations and for commercial formula- tribute to a lack of a positive effect on solubility or even
tions, for example, chlordiazepoxide (Librium; pro- a decrease in total drug solubility on the addition of
pylene glycol 20%, pH 3), D.H.E. 45 (15% glycerin, a cyclodextrin to a cosolvent formulation. First, the
6.1% ethanol, pH 3.6), fenoldopam (Corlopam; pro- cyclodextrin may complex with the cosolvent itself.
pylene glycol 50%, pH 3), oxytetracycline (Terramycin; Analogous to the effect of a surfactant, this leads to
propylene glycol 65–75%, pH 3), pentobarbital [Nem- a decrease in effective cosolvent concentration (He
butal; Lundbeck Inc., Deerfield, IL (propylene glycol et al., 2003), and simultaneous competition between
40%, ethanol 10%, pH 9.5)], and phenytoin (Dilantin; drug and cosolvent for the cyclodextrin hydrophobic
propylene glycol 40%, ethanol 10%, pH 10–12.3) (also cavity will reduce the complexation efficiency of the
see Section IV.B.2.a). cyclodextrin toward the drug (van Stam et al., 1996;
2. Cosolvents and Surfactants. Micellar solubiliza- Miyake et al., 1999b; Evans et al., 2000; Partyka et al.,
tion is a common mechanism by which the solubility 2001). The complexation efficiency may also decrease
Strategies for Low Drug Solubility 373

Fig. 24. Solubilization of NSC-639829 by ethanol, propylene glycol, DMSO, and PEG 400 at pH 1, 2, and 7. DMSO, dimethyl sulfoxide; PEG,
polyethylene glycol. Adapted from Jain et al. (2001).

in the presence of a cosolvent by virtue of the drug ethanol solutions is higher in the presence of hydrox-
showing more affinity for the bulk medium (which is ypropyl-b-cyclodextrin and that as little as 0.2%
less polar in the presence of a cosolvent) and, in turn, cosolvent (which is insufficient to solubilize a significant
a decreased tendency toward the formation of the drug- amount of drug via cosolvency alone) increases drug
cyclodextrin inclusion complex. In contrast, factors solubility in the presence of the cyclodextrin (Li et al.,
that may contribute to increased drug solubility in 1999d; He et al., 2003). Consistent with these data, the
cosolvent-cyclodextrin solutions include 1) higher con- currently marketed Sporanox IV solution (itraconazole;
centrations of free drug in the presence of a cosolvent, Janssen Pharmaceuticals, Titusville, NJ) contains 40%
providing an increased driver toward complexation; hydroxypropyl-b-cyclodextrin and 2.5% propylene gly-
and 2) the potential for the formation of highly soluble col (Strickley, 2004) (see also Section VIII.D.2).
ternary complexes of drug-cosolvent-cyclodextrin, par- 4. Cosolvents and Strong Electrolytes. Strong elec-
ticularly when the cosolvent and drug molecules are trolytes such as sodium chloride increase solution
small, as this may facilitate simultaneous drug and polarity, and this acts to oppose the decrease in po-
cosolvent complexation by the cyclodextrin (Reer and larity stimulated by addition of cosolvent. Conse-
Muller, 1993; Li et al., 1999d; Yalkowsky, 1999; Nandi quently, the addition of salts to cosolvent systems is
et al., 2003). The latter effect is exemplified by studies generally avoided as it will decrease drug solubility
showing that the solubility of fluasterone in aqueous (Yalkowsky, 1999).
374 Williams et al.

E. Effects of Dilution on Drug Solubilization of cosolvent formulations. Indeed, in an analysis of 21


by Cosolvents marketed formulations, Johnson et al. (2003) showed
Dilution of cosolvent solutions in aqueous media, that the effect of injection rate on drug precipitation
including biologic fluids such as blood, leads to a linear was drug/formulation dependent and that slowing the
decrease in drug concentration with increasing di- injection rate did not always reduce precipitation For
example, slowing the rate of infusion reduced pre-
lution. In contrast, the solubilizing capacity of cosol-
cipitation from Valium, Cordarone, nafcillin, and
vent formulations drops exponentially on dilution
phytonadione formulations, but the same approach
(green line, Fig. 25). In many cases, particularly where
led to increased precipitation from Dilantin and
cosolvent systems contain drug dissolved at concen-
dobutamine formulations. Similarly, using a dynamic
trations approaching the maximum solubility in the
precipitation model, Ward and Yalkowsky (1993)
system (e.g., the red line in Fig. 25), dilution has the
showed that it was possible to reduce amiodarone
potential to generate conditions where drug concen- precipitation by increasing (rather than reducing) the
trations are supersaturated (i.e., the required concen- rate of drug injection into the circulating medium. The
tration on dilution is greater than the drug solubility authors suggested that the use of a high infusion rate
in the system, X to Y in Fig. 25), and under these reduced the efficiency of mixing (i.e., the dilution rate)
circumstances, drug precipitation is likely. Drug pre- in blood, thereby reducing the risk of precipitation
cipitation after parenteral administration may cause (Ward and Yalkowsky, 1993). In vitro methods for
mechanical or chemical irritation at the injection site, assessing the likelihood of drug precipitation on
leading to phlebitis (Falchuk et al., 1985; Avis et al., dilution of solubilizing formulations are described in
1986), and potentially more serious systemic effects more detail in Section IV.B.3.a.
such as pulmonary embolism (Taniguchi et al., 1998).
Solid precipitates are also likely to cause pain during F. Potential Pharmacological Properties of Cosolvents
the administration of i.m. injections. 1. In Vitro Assessment of Cosolvent Toxicity. Pain on
Approaches to minimize the potential risk of pre- administration of cosolvent formulations is common
cipitation are therefore critical. This may be achieved and has been largely attributed to hemolysis of red
by 1) reducing the drug load in the formulation blood cells (Strickley, 2004; Amin and Dannenfelser,
(i.e., moving toward the blue line in Fig. 25 rather 2006). It is therefore routine to screen potential
than the red line); 2) reducing the rate of administra- cosolvent formulations for their potential toxicity to
tion (injection); 3) injecting into larger blood vessels red blood cells in vitro. In the method of Reed and
(allowing additional time and a greater supply of Yalkowsky (1985), 0.1 ml of formulation is vortex-
plasma proteins, lipoproteins, and red blood cells to mixed with various volumes of blood and incubated at
which drugs may bind); and 4) the use of combination 25°C for 2 min before mixing with 5 ml of saline to
formulations such as cosolvent-surfactant and cosolvent- quench the potential hemolytic properties of the co-
pH systems that are less susceptible to solubility de- solvent. The level of hemolysis is detected via an
creases on dilution. increase in optical density of the mixture (i.e., a change
These approaches, however, are not always effective of color resulting from the release of hemoglobin after
in eliminating the risk of drug precipitation on dilution cell lysis). In analogous dynamic tests, the formulation
is continuously infused into blood pumped at a given
rate before the mixture passes into a saline solution
to quench the hemolysis potential. A formulation is
considered compatible if it causes hemolysis in less
than 10% of cells, whereas formulations resulting in
25% hemolysis during a static test and 2% hemolysis
during the dynamic test are classified as hemolytic
(Amin and Dannenfelser, 2006).
Data obtained using hemolysis models are highly
dependent on experimental variables. For example,
levels of in vitro hemolysis are highly dependent on
the cell-cosolvent contact time and static methods typi-
cally result in higher levels of cell lysis compared with
results obtained from dynamic tests and in vivo studies
(Krzyzaniak et al., 1997). As such, very short (1 s) con-
Fig. 25. Effect of lowering cosolvent concentration through dilution on tact times and formulation-to-blood ratios of 0.1 have
the dissolved drug concentration. Points on the dilution curves at which been suggested to capture most adequately the potential
theoretical drug concentrations exceed maximum solubility (labeled X
and Y) indicate areas of precipitation risk. S0 is intrinsic drug solubility in toxic effect of a formulation. Traditionally, human blood
the absence of cosolvent. is used to assess the lytic properties of parenteral
Strategies for Low Drug Solubility 375

formulations, although rabbit blood has generated Cosolvents may be used for both neutral and ionizable
consistent cell lysis (Amin and Dannenfelser, 2006). compounds and have been used successfully in com-
2. Lytic Effects of Commonly Used Cosolvents. bination with other solubilization strategies, includ-
The potential for commonly used cosolvents to cause ing surfactants, cyclodextrins, pH manipulation, and
hemolysis has been widely reported (Mottu et al., lipids. The factors that should be considered when
2000). In a static cell lysis model, the concentrations (% using cosolvent formulations include the likely loss
v/v) of cosolvent required to induce lysis in 50% of blood of solubilization power on dilution (both before admin-
cells varied in the following order: dimethyl isosorbide istration or following mixing with blood) and the
(least toxic 5 39.5%) , DMA (37.0%) , PEG 400 cell-based toxicity of cosolvent, in particular, after
(30.0%) , ethanol (21.2%) , propylene glycol (5.7%) , i.v. administration.
DMSO (5.1%) (Reed and Yalkowsky, 1985). The same
rank order of hemolytic potential was reported in
a later study using similar methods (Fu et al., 1987). VII. Surfactants
Dynamic models have also suggested that propylene
glycol may be more toxic than ethanol (Krzyzaniak A. Rationale for the Use of Surfactants to Enhance the
et al., 1997). Extending these suggestions into in vivo Solubilization of Poorly Water-Soluble Drugs
studies in rats (Fu et al., 1987) and dogs (Fort et al., Surfactants are commonly used to solubilize poorly
1984), where animals were intravenously dosed vari- water-soluble drugs via drug incorporation into surfac-
ous cosolvent formulations on a daily basis over a 2- tant micelles, and they may also support enhanced
week period, ethanolic (10%) formulations containing delivery via improvements to wetting and in the
propylene glycol (40 and 50%) again showed more stabilization of suspension and nanosuspension for-
irritation and hemolysis (detected via increased hemo- mulations (Attwood and Florence, 1983; Yalkowsky,
globin content in the urine) compared with equivalent 1999; Malmsten, 2002). Micellar solution–based for-
ethanolic formulations containing up to 40% PEG 400. mulations are simple and effective and as such have
The toxic effects of propylene glycol toward red blood been widely used, largely after i.v. administration.
cells have been attributed to the hypotonicity of Some examples of marketed parenteral products include
propylene glycol solutions, although there is evidence paclitaxel (Taxol; Bristol-Myers Squibb), cyclosporine
that this may be attenuated both in vitro and in vivo (Sandimmune; Novartis AG, Basel Switzerland), amio-
when used in combination with ethanol (Reed and darone hydrochloride (Cordarone Intravenous; Pfizer),
Yalkowsky, 1985; Amin and Dannenfelser, 2006) or and calcitriol (Calcijex; Abbott Laboratories).
PEG 400 (Fu et al., 1987; Amin and Dannenfelser, Several surfactant-based parenteral formulations
2006) and in saline or isotonic phosphate (compared have been developed as nonaqueous preconcentrates
with water) (Reed and Yalkowsky, 1986; Fu et al., [docetaxel (Taxotere; Sanofi-Aventis, Bridgewater, NJ),
1987; Amin and Dannenfelser, 2006). Indeed, it should Prograf, Taxol) and therefore require significant di-
be noted that propylene glycol and ethanol are by far lution before use. Sandimmune, for example, comprises
the most widely used cosolvents in parenteral formu- drug dissolved in a combination of surfactant and cosol-
lations. Ativan injection (containing lorazepam), for vent (65% Cremophor EL, 35% ethanol) that is diluted
example, contains 80% propylene glycol and is diluted at least 50- to 100-fold with normal saline or 5%
only 1:1 with saline, dextrose 5%, or lactated Ringer’s dextrose before i.v. administration. The requirement for
solution before administration via an i.v. or i.m. bolus dilution before administration, and the inevitable di-
(Strickley, 2004). The strong lytic properties of DMSO lution that occurs postinjection, leads to the possibility
limits its use during preclinical and clinical testing (Li of drug precipitation; to avoid this, a number of combi-
and Zhao, 2007), although it has been used in pre- nation surfactant-containing formulation approaches
clinical studies at concentrations below 20%. The sys- have been explored (Li et al., 1999c; Li and Zhao,
temic effects of commonly used cosolvents have been 2003; Jain et al., 2004; Strickley, 2004; Li and Zhao,
extensively described by Buggins et al. (2007) and 2007). These include surfactant-cosolvent combinations
Mottu et al. (2000). Recently, Gad and coworkers and surfactants combined with pH-adjustment strate-
(2006) performed an extensive review of 15 years of gies. These approaches reflect the realization that a
in vivo preclinical studies in which the safety of the single solubilization strategy is often insufficient to
formulation vehicle was assessed. Table 17 lists cosol- solubilize the entire dose of drug. Combination appro-
vent regimens after single or repeat dosing in various aches allow the use of lower concentrations of a single
animals that were all “well tolerated.” component, a strategy that has found favor since high
surfactant concentrations have the potential to cause
local or systemic adverse reactions after parenteral ad-
G. Summary
ministration, including pain at the injection site or more
Cosolvents are widely used to solubilize poorly water- serious hypersensitivity reactions (Gelderblom et al.,
soluble drugs after oral and parenteral administration. 2001; ten Tije et al., 2003; Strickley, 2004).
376 Williams et al.

TABLE 17
Cosolvent regimens “well tolerated” in various animals
Summarized in this table are the data obtained in a number of in vivo studies performed by four separate organizations between 1991 and 2006 and summarized in Gad
et al. (2006). In each study, a vehicle control group was evaluated, with the results summarized in this table.

Cosolvent Species Duration of Study Route of Administration Dose

DMSO Dog 1 mo i.v. 1.25 ml/kg


Rat 1 mo i.v. 200 mg/kg
Rat 1 mo i.p. 5 ml/kg
Rat 1 mo Oral 5 ml/kg
Guinea pig 1 mo i.v. 0.1 mg/kg
Mouse 1 mo Oral 5 ml/kg
Rabbit 1 mo s.c. 1 ml/kg
Primate “Efficacity” Oral 3 ml/kg/day
Ethanol Dog 1 mo Oral 5 ml/kg (7.5% solution)
Dog 3 mo Oral 5 ml/kg (5.0% solution)
Rat 1 wk Oral 10 ml/kg (10% solution)
Mouse 6 mo Oral 2.5 g/day
Mouse 1 mo Oral 2.5 ml/day (5% solution)
Mouse Acute i.p. 5 ml/kg (5% solution)
Mouse 13 wk Cutaneous 100 ml/animal/day (70% solution)
Glycerol Rat 1 mo Oral 1000 mg/kg
Rat — s.c. 10 mg/kg
Guinea pig 1 mo Oral 500 mg/kg
Mouse 1 mo i.v. 100 mg/kg
Mouse Acute s.c. 10 mg/kg
Mouse 1 mo i.p. 10 mg/kg
Rabbit — i.v. 10 mg/kg
PEG 300 Guinea pig 1 mo i.v. 1 ml/kg
Mouse ADME test Oral 10 ml/kg/day
Rabbit 1 mo Oral 500 mg/kg
PEG 400 Guinea pig 1 mo Oral 1000 mg/kg
Mouse 1 mo Oral 10 ml/kg/day
Mouse 1 mo i.p. 500 mg/kg
Rat 1 mo Oral 5 ml/kg
Rat 1 mo i.v. 0.5 ml/kg
Rat 1 mo Cutaneous 5 ml/kg (35% solution)
Rat 104 wk Cutaneous 2.5 ml/kg
Minipig 2 wk Cutaneous 2.5 ml/kg
PG Rat 1 mo Oral 2.5 ml/kg
Rat 1 mo s.c. 2.5 ml/kg
Minipig 26 wk Cutaneous 2.5 ml/kg
Mouse 1 mo Oral 10 ml/kg (50% solution)
Mouse 1 mo i.p. 2.5 ml/kg (40% solution)
Dog 1 mo Oral 2.5 ml/kg
ADME, absorption, disposition, metabolism, and elimination; PG, propylene glycol; mo, month; wk, week.

Unlike most drug solubilization technologies, where combination approaches that include surfactant, cosol-
in vivo pharmacological effects of the formulation com- vent, and lipid. Such formulations are usually de-
ponents are limited, in the case of surfactants in- scribed under the generic heading of “lipid-based
creasing evidence suggests the potential for inhibitory formulations”, regardless of the quantity of tradi-
effects on both metabolic and drug transport or anti- tional “lipid” incorporated. Notable examples of for-
transport processes (Martin-Facklam et al., 2002b; mulations of this type include amprenavir (Agenerase;
Singla et al., 2002; Bogman et al., 2003). In particular, GlaxoSmithKline), cyclosporine (Neoral; Novartis
inhibition of cellular efflux processes might be expected Pharmaceuticals), and ritonavir (Norvir SEC; Abbott
to enhance drug absorption for compounds where Laboratories). These systems are described in more
intestinal efflux transporters such as P-glycoprotein detail in Section XI of this review. Surfactants are also
(P-gp) limit oral bioavailability or provide enhanced commonly used to stabilize nanosuspension formula-
cellular access to, for example, multidrug-resistant tions and are therefore an important constituent of
tumor populations where resistance is mediated by formulations described in Section IX. Finally, surfac-
overexpression of P-gp. Surfactant inhibition of P-gp tants may be used to improve wetting of hydrophobic
activity at the blood-brain barrier has also been sug- materials in simple solid dosage forms and to stabilize
gested to promote drug access to the central nervous traditional suspension-based formulations. The inter-
system (Kabanov et al., 2003). ested reader is directed to the following reviews for
In products for oral administration, surfactants are further details (Nielloud and Marti-Mestres, 2000;
rarely used as the only solubilizing agent, either in Tadros, 2005).
liquid formulations or in solubilizing formulations de- The emphasis of the current section is on the use of
stined for filling into soft or hard gelatin capsules. For surfactants in parenteral formulations. In comparison
these applications, surfactants are typically used in with oral delivery (where most surfactants are not
Strategies for Low Drug Solubility 377

absorbed), this introduces additional safety concerns surfactants that are frequently used in pharmaceutical
since injected formulations are directly introduced into formulations. Nonionic surfactants are generally esters
the systemic circulation and rapidly interact with of saturated, unsaturated fatty acids or glycerides,
vascular tissue after administration. Here we describe monoethers of fatty alcohols, or block copolymers
the surfactants that are commonly used in parenteral (polyethers) of polyethylene oxide and polypropylene
formulations, introduce the theory underpinning drug oxide. They may be grouped according to the nature of
solubilization by micellar systems, address the formu- the hydrophilic head group, which includes sorbitan
lation and nonformulation factors that may affect (used in Spans) and its ethoxylated derivatives (used in
solubilization, and finally, and discuss the potential Tweens) or polyoxythelene chains of varying length
pharmacological effects of surfactants. (present in surfactants such as Cremophor, Gelucire,
Brij, Labrasol, and Labrafil). Surfactants can be manu-
B. Commonly Used Nonionic Surfactants in factured via ethoxylation (e.g., reaction of sorbitan esters
Drug Delivery with ethylene oxide in the case of Tween surfactants) or
The surface activity of surfactants is derived from esterification (e.g., esterification of medium-chain-length
the presence of both hydrophilic and lipophilic environ- fatty acids in the case of Labrasol). Surfactants obtained
ments in the molecular architecture. Depending on the by ethoxylation may show higher purity, although
charge on the hydrophilic segment (often referred to as molecular comparison of surfactants manufactured via
the “head group,” as opposed to the lipophilic “tail”), different processes is complex (Schick, 1966).
surfactants may be categorized as either ionic (anionic The poloxamer block copolymers (Pluronics, Lutrols,
or cationic) or nonionic. Nonionic surfactants are more Synperonics) comprise a hydrophobic poly (propylene
common in drug delivery owing to their better safety, oxide) central segment flanked on either side by more
superior capacity to solubilize nonpolar solutes at lo- hydrophilic ethylene oxide blocks (Kabanov et al.,
wer concentrations, and good compatibility with other 2002) and are a class of nonionic polymeric surfactants
pharmaceutically relevant excipients (Elworthy and that have attained significant recent interest for drug
Treon, 1966; Davis et al., 1970; Nielloud and Marti- solubilization and targeted drug release (Adams et al.,
Mestres, 2000). Table 18 lists a range of nonionic 2003; Chiappetta and Sosnik, 2007).

TABLE 18
Examples of nonionic surfactants commonly used as solubilizing agents in parenteral drug delivery

General Class Primary Hydrophilic Primary Hydrophobic Component Cloud Point


Component

°C
Span Sorbitan Monolaurate (Span 20) LS
Monopalmitate (Span 40)
Monostearate (Span 60)
Monooleate (Span 80)
Sesquioleate (Span 83)
Trioleate (Span 85)
Tween Polyethoxylated sorbitan Monolaurate (Tween 20) 76
(polysorbate) Monopalmitate (Tween 40) 73
Monostearate (Tween 60) 80
Monooleate (Tween 80) 65
Trioleate (Tween 85) LS
Cremophora Polyoxyethylene Hydrogenated castor oil (Cremophor RH40) 93
Castor oil (Cremophor EL) 72
b
Polyoxylglycerides Polyoxyethylene Caprylic and capric glycerides (Labrasol) LS
Oleyl or linoleyl glycerides (Labrafil M1944CS, LS
M2125CS)
Stearoyl or lauroyl glycerides (Gelucire 50/13, 44/14) 64
Tocols Polyethylene glycol D-a-Tocopherol succinate ;40
Brij Polyoxyethylene Lauryl alcohol (Brij 35) .100
Cetyl alcohol (Brij 58) .100
Stearyl alcohol (Brij 78)
Oleyl alcohol (Brij 98)

Poloxamers Ethylene Propylene


oxide units oxide units

153 23 Poloxamer 188 (Pluronic F68) .100


27 23 Poloxamer 184 (Pluronic L64) 58
8 43 Poloxamer 331 (Pluronic L101) 15
74 43 Poloxamer 335 (Pluronic P105 91
LS, low aqueous solubility limits accurate determination of cloud point.
a
Ethoxylated castor oil derivatives contain a composite mix of monoesters, diesters and triesters of polyethoxylated fatty acids (primarily ricinoleic acid) and hydrophilic,
nonfatty acid esterified materials, including glycerol, polyoxyethylene ether, and free polyethylene glycol.
b
Also known as macrogolglycerides and are composed of a well defined mixture of monoglycerides, diglycerides, and triglycerides, monoesters and diesters of PEG, and
some free PEG.
378 Williams et al.

Surfactants are also commonly classified by their 1. Micelle Formation and Drug Solubilization.
hydrophilic-lipophilic balance (frequently abbreviated The amphiphilic nature of surfactants promotes self-
to HLB), which categorizes surfactants on the basis of association in aqueous environments and facilitates
the molecular weight (MW) ratio of their hydrophilic the formation of micelles above a critical concentration,
and lipophilic components, expressed as a percentage frequently referred to as the critical micelle concentra-
divided by five (Griffin, 1949). For traditional nonionic tion (CMC). Surfactant micelles have the capacity to
surfactants, HLB values range from 0 (most lipophilic) enhance the effective aqueous solubility of poorly
to 20 (most hydrophilic); for poloxamers, the HLB water-soluble drugs, largely via solubilization in the
value is determined by the ratio of the lengths of hydrophobic micelle core but also via interaction with
ethylene oxide and propylene oxide blocks. Apparent head groups and orientation/incorporation into the
HLB values for Pluronics and other block copolymers polar/nonpolar water-micelle interface. The location of
can exceed 20 (Kozlov et al., 2000). solubilization is largely dependent on drug hydropho-
The highest concentrations of several nonionic bicity and charge (Mukerjee, 1979; Yalkowsky, 1999;
surfactants that have been used in commercial paren- Rangel-Yagui et al., 2005). Highly lipophilic drugs are
teral formulations are presented in Table 19 and expected to be primarily solubilized by lipophilic sur-
provide an indication of the maximum quantities that factant “tails” (e.g., alkyl chains) orientated to the
might be tolerated in human subjects. Higher concen- micelle core, whereas less lipophilic compounds are
trations may be possible during preclinical testing, predominantly solubilized at the polyoxyethylene-rich
especially over short periods, although the potential core/mantle interface (Elworthy and Patel, 1982;
pharmacological effects of surfactant must be consid- Fahelelbom et al., 1993; Malcolmson et al., 1998).
ered when developing formulations for toxicology test- The CMC reflects the saturated solubility of surfac-
ing. Concentrations of administered surfactant are tant monomers in the solvent (e.g., water) at a partic-
often higher if the formulation is to be given via an i.m. ular temperature (Becher, 1967). Micelle formation is
injection, although lower injection volumes must be driven thermodynamically by the liberation of struc-
used (Nema and Brendel, 2011). tured water that is involved in the hydration of the
hydrophobic alkyl chain in monomeric solution, back
C. Solubilization by Surfactants into bulk solution. Micelle formation also results in
Effective micellar solubilization of poorly water- removal of hydrophobic regions of the surfactant mol-
soluble drugs is dependent on several factors, including ecule from an aqueous environment (when surfactant
surfactant properties, drug properties, and a variety of monomers are present in free solution) and incorpora-
external factors such as the presence of other solubi- tion into the largely hydrophobic micellar core. This
lizing agents, electrolytes, and dilution. An under- results in a gain in entropy through increased mobility
standing of the factors that determine solubilization of the alkyl chain in the micelle core and reduction in
efficiency is critical to the rational design of micellar water structure.
formulations and allows early identification of circum- The higher aqueous solubility of ionic surfactants
stances where alternative solubilization strategies and their potential to repel adjacent surfactant mol-
might be more appropriate (Li et al., 1999a,c). ecules via electrostatic repulsion at the micellar

TABLE 19
Commonly used surfactants: published recommendations for concentrations in preclinical testing and maximum concentrations observed in clinical
parenteral products
Recommended Concentrations Used in
Surfactant Range for Frequency of Use in Some Commercial Final Surfactant Concentration Administered
Preclinical Usea Commercial Formulationsb Preconcentratesc
(oral and i.v.)

Cremophor EL 5–10% 3 65% (Sandimmune)d #5% after a 20- to 100-fold dilution, i.v. infusion.
Average volume per single dose 5 3.5 mle
51% (Taxol)d #10% after 5- to 20-fold dilution, i.v. infusion.
Average volume per single dose 5 25.8 mle
50% (Vumon) #5% after a 10- to 100-fold dilution, IV infusion
50% (Valstar) 18% after a ;4-fold dilution, intravesical instillation
Cremophor RH 40 5–10% 1 7%f —
Cremophor RH 60 1 20% (Prograf) #1% after a 250- to 1000-fold dilution, i.v. infusion
Tween 20 22 2.4%f
Tween 80 5–10% 72 100% (Taxotere)d #2% after a ;50- to 150-fold dilution, i.v. infusion
Solutol HS-15 20–50% 50% (Panitol)g
a
% w/v on the basis of single 10 ml/kg excipient dose to a mouse or rat, from Li and Zhao (2007).
b
From Nema and Brendel (2011).
c
From Strickley (2004).
d
Significant evidence of local/systemic adverse reactions.
e
From Gelderblom et al. (2001).
f
It was not possible to identify the product.
g
Licensed for use in Mexico.
Strategies for Low Drug Solubility 379

interface results in a higher CMC in comparison with formulations where dilution leads to a nonlinear de-
nonionic surfactants (with similar hydrophobic seg- crease in drug solubility and, typically, drug precipi-
ments) (Malmsten, 2002). At equal total surfactant tation. The molar solubilization capacity k is frequently
concentration, therefore, a lower concentration of an used to quantify surfactant solubilization and is
ionic surfactant will be present as micelles. defined as the mole ratio of micellar solubilized drug
Several physicochemical properties of surfactant (SM) to the concentration of micellar surfactant (Cmic)
solutions change dramatically at the CMC, and, (Yalkowsky, 1999):
therefore, CMC measurement is readily achieved via ðStot 2 S0 Þ S
a change in surface tension, light scattering, molecular k5 5 M ð35Þ
Csurf 2 CMC Cmic
conductivity, osmotic pressure, self-diffusion, and mag-
netic resonance with increasing surfactant concentra- where S0 is the intrinsic solubility of the drug, Stot is
tion. The physical properties, including CMC, for a the total concentration of drug in solution (solubilized
range of ionic and nonionic surfactants are well de- plus free), and Csurf is the total surfactant concentra-
scribed in detail elsewhere (Elworthy et al., 1968;
tion. The determination of k provides a rapid, effective
Yalkowsky, 1999); cloud-point data (providing a mea-
means of comparing the solubilization efficiency of
sure of the aqueous solubility of a surfactant) for a
various surfactants for a particular drug. It does not,
number of commonly used nonionic surfactants are
however, account for potential differences in the in-
included in Table 18. Surfactants typically form
trinsic drug solubility. The micelle-water partition
spherical micelles in water, as this conformation
coefficient (KM) is therefore a more appropriate index
maximizes the surface area of interaction of the polar
for comparing total solubilization for different drugs:
head groups with the external aqueous environment
and minimizes the aqueous interaction of the nonpolar ðStot 2 S0 Þ SM
tail (Tanford, 1974, 1980). For a long-chain nonionic KM 5 5 ð36Þ
S0 S0
surfactant, the number of surfactant monomers within
a spherical micelle may be as many as 150 (Arnarson
and Elworthy, 1982), although further growth of a The aforementioned solubilization descriptors have
spherical micelle is usually limited by steric con- been used widely to quantify and compare drug solu-
straints (Attwood and Florence, 1983). Certain amphi- bilization by various surfactants and to explore the
philes adopt nonspherical micellar conformations, impact of drug physicochemical properties on solubiliza-
typically rod-shaped micelles, and these may further tion (Yalkowsky, 1999; Alvarez-Nunez and Yalkowsky,
aggregate to form ordered hexagonal, cubic, or lamellar 2000; Croy and Kwon, 2004; He et al., 2006; Kadam
liquid crystal phases (Attwood and Florence, 1983; et al., 2009; Kadam et al., 2011). For example, using
Laughlin, 1994; Lawrence, 1994). Dispersed, self- Tween 80 as a model surfactant, Alvarez and co-
assembled liquid crystals comprising amphiphilic oils workers showed that increasing the octanol-water
(e.g., phospholipid, monoolein) alone or in combination
with other surfactants have been examined as solubil-
izers, most commonly liposomes (dispersed lamellar
phase), but also cubosomes and hexasomes (Lawrence,
1994; Boyd et al., 2009). Liposomes, cubosomes, and
hexasomes may be used to overcome low drug solu-
bility in i.v. drug formulations; however, their complex
structure commonly alters the pharmacokinetics and
pharmacodynamics of encapsulated drug molecules.
As such, they have found greater application in the
development of targeted drug delivery systems with
controlled drug release profiles. Here we focus on
simple micellar systems for drug solubilization; the
interested reader is directed to alternate texts de-
scribing the use of liquid crystalline systems for drug
delivery (Lawrence, 1994; Shah et al., 2001; Smart
et al., 2008; Boyd et al., 2009).
2. Quantification of Micellar Solubilization. For
poorly water-soluble drugs, increasing surfactant con-
centrations above the CMC results in linear increases Fig. 26. Solubility of dexamethasone in water (S0) and as a function of
in drug solubility (Fig. 26). Where the CMC is low, this surfactant aqueous concentrations (% w/w) of long-chain nonionic
surfactants. OEXX denotes the number of oxyethylene units. S0 is the
also provides for a solubilization profile that is largely intrinsic drug solubility in the absence of surfactant. Adapted from Barry
linear on dilution, a situation in contrast to cosolvent and El Eini (1976).
380 Williams et al.

partition coefficient (log P) of a solubilized drug re- whereas solubilization of the more lipophilic vitamin A
sulted in a corresponding increase in KM (Alvarez- palmitate is higher in micelles constructed with the
Nunez and Yalkowsky, 2000) (Fig. 27) and led to the longer-chain Tween 80 surfactant (Yalkowsky, 1999).
suggestion that micellar solubilization is likely to be of The effect of surfactant chain length also diminishes
most potential benefit for drugs with log P values . 3. above a chain length of 14–16 carbons atoms (Elworthy
The exceptions to this suggestion include drugs with and Patel, 1982; Ong and Manoukian, 1988; Yalkowsky,
high potency that require relatively low levels of solu- 1999) (Fig. 28). In general, the lower CMC values of
bilization and drugs where interaction at the micellar nonionic versus ionic surfactants promote more effec-
interface is the prevalent mechanism of solubilization. tive drug solubilization, although this is not always
3. Effect of Surfactant Type and Structure on Drug the case, and examples of higher drug solubilization
Solubilization. Increasing the length of the hydro- with ionic surfactants have been documented (Krishna
phobic region of an ionic or nonionic surfactant leads and Flanagan, 1989; Bhat et al., 2006; Bhat et al.,
to a decrease in the water solubility of surfactant 2008; Sanghvi et al., 2009). In these circumstances, it is
monomers, in turn leading to a decrease in the CMC likely that solubilization within the micelle-water in-
(Hall and Pethica, 1967; Attwood and Florence, 1983; terface or direct interaction with ionic head groups
Chandler, 2005). Micelle formation from surfactants dominates interaction patterns rather than solubiliza-
with larger hydrophobic chains also typically results in tion within the micelle core (Malcolmson et al., 1998).
an increase in the volume of the micelle core (El-Eini Where interfacial interactions dominate, the molecular
et al., 1976; Lindman et al., 1980; Ong and Manoukian, weight of the polyoxythelene component of the surfactant
1988; Hurter et al., 1993; Wanka et al., 1994; Chen may prove to be the most significant factor in deter-
et al., 1998; Tsui et al., 2010). Together, these trends mining solubilization properties.
predict an improvement in micellar drug solubilization 4. Effect of Temperature on Drug Solubilization.
with surfactants of increasing alkyl chain length. The Micellar formulations may be subjected to elevated
effect of surfactant alkyl chain length has been ob- temperatures on storage or during autoclaving to
served in many solubilization studies using Tween, provide for sterility following manufacture (Schott
Brij, and Pluronic surfactants (Ismail et al., 1970; and Han, 1975; Na et al., 1999). Conversely, low temp-
eratures may be encountered where refrigeration is
Samaha and Cadalla, 1987; Ong and Manoukian, 1988;
required to provide appropriate drug stability. An
Kozlov et al., 2000; Ahmed, 2001; Tommasini et al.,
understanding of the potential impact of temperature
2004; Kovacs et al., 2009).
on micellar solubilization is therefore important during
Other reports suggest that the impact of surfactant
formulation design.
chain length on drug solubilization is significant only
At a fixed surfactant concentration, increasing
when the drug is highly nonpolar and has a high
temperature results in a reduction in the degree of
affinity for the micelle core (Alvarez-Nunez and
Yalkowsky, 2000). For example, indomethacin solubi-
lization is similar in Tween 20 or Tween 80 micelles,

Fig. 28. Solubility enhancing effect of Tween 20, 60, and 80 at


Fig. 27. Logarithm of Tween 80 molar micelle-water partition coefficient concentrations of 1–3% on the aqueous solubility of miconazole. The
log KNM, versus log P profile of tested drugs. Note that log KNM is the log figure shows that increased drug solubilities are obtained on increasing
KM value normalized to a 1.0 M surfactant concentration. Log P is the the alkyl chain length of Tween surfactants. Solubility differences are
octanol-water partition coefficient. Adapted from Alvarez-Nunez and most pronounced between Tween 20 and Tween 60/Tween 80. Adapted
Yalkowsky (2000). from Kovacs et al. (2009).
Strategies for Low Drug Solubility 381

hydration of the polyoxyethylene chains of nonionic therefore, their increased ability to alter the polarity of
surfactant monomers (Schott and Han, 1975; Na et al., the solvent compared with ethanol (Kawakami et al.,
1999). Ultimately, this leads to phase separation of the 2006). Interestingly, alcohols of more than four carbon
surfactant as the solubility limit is broached. This is atoms have the opposite effect, causing a decrease in
the cloud point, that is, the point at which an increase the CMC of nonionic surfactants (Armstrong et al.,
in temperature results in the generation of a cloudy 1996; Bharatiya et al., 2007).
solution (usually defined as an increase in solution To consider the net effect of cosolvents on drug
turbidity of 50%). As a result, the CMC for nonionic solubilization in a micellar formulation requires the
surfactants (the opposite is true for ionic surfactants) appreciation of a number of factors. As described
decreases on heating (Wanka et al., 1994; Tadros, already, cosolvents usually lead to an increase in
2005), which increases the potential for drug solubili- surfactant CMC, but they also increase drug solubility
zation. However, temperatures in excess of the cloud in the bulk phase, causing drug partitioning into the
point are likely to lead to formulation instability. The micelle to decrease. This may be further complicated if
effect of changes to temperature on drug solubilization the cosolvent itself is solubilized by the surfactant as
in micellar systems is complex since temperature also the micelle core will become more polar and, therefore,
impacts on the intrinsic aqueous solubility of the drug. less attractive to drug solubilization (Zana, 1995;
In most cases, a drug will show greater affinity for the Kawakami et al., 2006). The combined use of cosolvents
bulk solution with increasing temperature and, in with surfactants may therefore lead to a decrease in
turn, a lower driving force for partitioning into the micellar drug solubilization, particularly when the
micellar phase. Although this suggests a possible de- drug shows high affinity for the surfactant core
crease in micellar solubilization, the net effect under (Kawakami et al., 2004; Dong et al., 2008). In contrast
conditions of increasing temperature is usually an to the potentially detrimental effect of cosolvents on
increase in overall solubility in nonionic surfactant micellar solubilization, cosolvency increases drug sol-
solutions (Humphreys and Rhodes, 1968; Barry and ubility in bulk solution. As such, the net impact on
Eleini, 1976; Yalkowsky, 1999). At lower temperatures, drug solubilization is difficult to predict a priori and
drug solubility in solution is reduced and the CMC of reflects the relative efficiencies of cosolvency versus
nonionic surfactants is expected to increase (as sur- micellar solubilization on drug solubility.
factant solubility increases). Decreasing temperature Surfactants are also commonly added to cosolvent
therefore increases the risk of drug precipitation. systems in an attempt to reduce susceptibility to
precipitation on dilution (see Section VI.D.2) (Tonsberg
D. Impact of Cosolubilizers and Electrolytes on et al., 2011) and are found in a number of commercial
Surfactant-Mediated Drug Solubilization parenteral formulations (Ni et al., 2002; Strickley,
1. Surfactants and Cosolvents. Cosolvency aims to 2004), including Taxol and Sandimmune. Both for-
reduce the polarity of an aqueous medium to improve mulations are nonaqueous solutions of Cremophor EL
the solvation of a hydrophobic drug and is described in and dehydrated alcohol, and both require dilution with
more detail in Section VI. Cosolvents may also be used significant quantities of aqueous phase before i.v.
in combination with surfactants. In this case, cosol- administration. The presence of cosolvent aids in
vents impact not only the solution properties of the dispersion and dilution of the surfactant solution.
drug but also the efficiency of micellar solubilization. 2. Surfactants and pH Adjustment. The properties
Thus, by reducing the polarity of the solution, of micelles formed by nonionic surfactants are largely
cosolvents enhance the solvation of nonionic surfac- independent of pH. However, changing pH can alter
tants, giving rise to increases in CMC and decreases in the micellar solubilization of ionizable drugs.
micelle size. Consistent with this suggestion, a wide For ionizable compounds, the total quantity of
range of pharmaceutical cosolvents (e.g., ethanol, dissolved drug is the sum of the concentrations of
propylene glycol, polyethylene glycol, DMA, and glyc- ionized [Di] and un-ionized [Du] drug in free solution
erol) have been shown to increase the CMC of aqueous and the concentrations of ionized [DiM] and un-
solutions of polyoxyethylene surfactants (Eicke, 1980; ionized drug [DuM] in the micellar phase. The
Yalkowsky, 1999; Kawakami et al., 2004; D’Errico micelle-water partition coefficients for ionized drug
et al., 2005) and the block copolymer surfactants (Na (Ki) are predictably lower than the corresponding
et al., 1999; Ivanova et al., 2001; Kawakami et al., values for the un-ionized form (Ku) as a result of
2006; Tomsic et al., 2006; Sarkar et al., 2010). For differences in solubility in the aqueous solvent (Jain
example, the addition of 10% ethanol to a micellar et al., 1999, 2001; Li and Zhao, 2003; He et al., 2006).
solution of Tween 80 led to an ;4-fold increase in Under conditions of differing pH, the concentration of
CMC, and the same concentration of PEG and DMA led un-ionized drug in solution (Du) remains unchanged,
to even greater (;10-fold and ;18-fold, respectively) and assuming no change to the properties of the
increases. The larger effect of PEG and DMA was nonionic surfactant micelle with pH, then the affinity
suggested to reflect their greater hydrophobicity and, of the un-ionized form for the micelle phase is also
382 Williams et al.

unchanged. Ku and Du are therefore pH independent and the presence of a nonionic surfactant has been
(Li et al., 1999b,c). achieved in several commercial micellar formulations,
In contrast, changing pH will alter the concentration including Cordarone (10% Tween 80, pH 4.1) and
of ionized drug in free solution [Di] and, depending on Librium (4% Tween 80, pH 3) (see also Section IV.B.2.c).
the magnitude of Ki (which describes the affinity of the The effect of pH change on micellar drug solubiliza-
charged drug for surfactant micelles), may increase the tion is more complex for ionic surfactants since pH
quantity of micellar ionized drug, [DiM]. Two extremes change alters the ionization state of both the drug and
might be envisaged. First, where the affinity of ionized the head group of the surfactant. For an anionic sur-
drug for micelles is low, the slope of drug solubility vs. factant, an increase in pH is expected to lead to greater
surfactant concentration plots at varying pH will be ionization of the head group and an increase in the
unaffected (and reflect KM, eq. 36) (Yalkowsky, 1999) CMC through the combined effect of increased solu-
but will be shifted upward, reflecting the higher bility and electrostatic repulsion between adjacent
solubility of ionized drug in free solution (Ran et al., surfactant head groups at the micellar interface. Un-
2005). Second, where ionized drug has significant af- der these circumstances, a decrease in drug solubili-
finity for micelles, the slope of the drug solubility zation is expected for nonionizable or anionic drugs
versus surfactant concentration plots will change with (where electrostatic repulsion from micellar head groups
a change in pH, reflecting contributions from both may reduce solubility further). In contrast, the poten-
micellar ionized and un-ionized drug. tial to form ion pairs may increase drug solubilization
The latter scenario is exemplified in Fig. 29, where for a weak base. Indeed, the rank order of surfactant
the increase in solubility of flavopiridol, a poorly sol- solubilization efficiency for a weak base (under con-
uble weak base (pKa 5 5.68), is shown as a function of ditions where both drug and anionic/cationic surfactant
pH and Tween 20 concentrations (Li et al., 1999b). are ionized) has been suggested to be anionic surfac-
In this example, the micellar association constant for tant . nonionic surfactant . cationic surfactant (Bhat
un-ionized flavopiridol (Ku) is 333.3 M21 and for et al., 2008). The opposite trend is expected for a weakly
ionized drug is 185.4 M21. Under conditions of de- acidic compound.
creasing pH, an increase in [Di] is therefore apparent; 3. Surfactant Mixtures. A combination of surfac-
in addition, micellar solubilization of the increasing tants may be used to promote drug solubilization
quantity of the ionized form is also enhanced, and as where the required concentration for a single surfac-
tant is high and likely to give rise to adverse effects.
a consequence, the increase in drug solubility with
The CMC of a binary mixture of two surfactants can be
increasing surfactant concentration is greater than
predicted using eq. 37, where x1 and x2 are the relative
that observed when only un-ionized drug is available in
mole fractions of the different surfactants in micellar
solution (i.e., at higher pH) (Li et al., 1999b). Similar
form (Tadros, 2005):
trends in solubility with changing pH and surfactant
concentration have been observed by others (Ikeda CMC 5 x1 CMC1 1 x2 CMC2 ð37Þ
et al., 1977; Jain et al., 2001, 2004; Ran et al., 2005;
In accordance with eq. 37, the CMC of a surfactant
Sheng et al., 2006; Kovacs et al., 2009). The potentially
mixture will be an average of the CMC of the single
beneficial effect of the combined use of pH adjustment
components. However, this relationship applies only to
systems where the two surfactants are structurally
related (i.e., of the same class with similar polar head
groups but different alkyl chain lengths).
In practice, it is more common to investigate sys-
tems containing two or more structurally unrelated
surfactants since their combined use more commonly
leads to the formation of mixed micelles with drug
solubilization properties that are greater than the
simple additive effect of the two surfactants. A classic
example of this type of synergy is the formation of bile
salt-phospholipid mixed micelles, which show sub-
stantial increases in solubilization capacity when used
in combination. Evidence of mixed micelles in solutions
containing ionic and nonionic surfactants has also been
widely reported (Treiner, 1994; Bhat et al., 2008), and
the combination of hydrophilic and hydrophobic Plur-
onic surfactants has been explored to obtain micelles of
Fig. 29. Experimentally determined total aqueous flavopiridol solubil-
ities (mM) in Tween 20 solutions at different pH values. Adapted from Li improved size, stability, and solubilization capacity
et al. (1999b). (Danson et al., 2004; Oh et al., 2004; Wei et al., 2009;
Strategies for Low Drug Solubility 383

Yang et al., 2009). Surfactant combinations are impact of electrolyte addition on drug solubilization in
widely used as components in lipid-based formula- ionic and nonionic micellar solutions has been explored
tions and are described in more detail in Section XI. using a range of sulfonylurea-type compounds and
4. Surfactants and Cyclodextrins. Cyclodextrins— glitazones (Seedher and Kanojia, 2009). These studies
including hydroxypropyl-b-cyclodextrin and sulfobutylether- report a greater than 10-fold increase in the solubili-
b-cyclodextrin—increase the apparent drug solubility of zation capacity of mixed micelles formed between ani-
a number of poorly water-soluble drugs by forming onic (i.e., sodium lauryl sulfate) and nonionic (Tween
drug-cyclodextrin inclusion complexes (see Section VIII). 80) surfactants toward glyburide, pioglitazone, and
Cyclodextrins can also complex with both nonionic rosiglitazone on the addition of 0.15 M sodium chloride
(Veiga and Ahsan, 1998; Buschmann et al., 2000; Eli (Seedher and Kanojia, 2009). Strong electrolytes have
et al., 2000) and ionic surfactants (Junquera et al., 1993; also been shown to depress the CMC of Pluronic block
Park and Song, 1989). As a result, the drug-cyclodextrin copolymers (Bahadur et al., 1992; Jain et al., 1999,
binding constant (a descriptor of drug solubilization by 2000; Desai et al., 2001).
cyclodextrins) is typically lower in the presence of
surfactant as a result of direct competition for available E. Effects of Dilution on Drug Solubilization
cyclodextrin. Furthermore, the complexed surfactant by Surfactants
fraction is unable to participate in micelle formation, Parenteral and oral micellar formulations are di-
leading to a reduction in micellar solubilization of poorly luted in situ after administration, and formulations for
water-soluble drugs (Junquera et al., 1993). Drug i.v. use are frequently diluted before administration.
solubilization in a solution containing both a surfactant Understanding the potential for loss of solubilization
and cyclodextrin is therefore unfavorable, and their capacity on dilution is therefore an important aspect of
combined use commonly provides no additional benefit dose form design for micellar formulations. As illus-
to total drug solubilization. Indeed, in some cases, trated in Fig. 30, the dilution of a micellar formulation
surfactant-cyclodextrin combinations lead to a decrease will typically only cause drug precipitation when the
in the concentration of solubilized drug (Veiga and concentrations of surfactant approach the CMC
Ahsan, 1998; Yang et al., 2004; Rao et al., 2006) (see also (marked in Fig. 30) and, in particular, when the
Section VIII.D.3). micellar formulation contains a level of drug (or drug
5. Surfactants and Strong and Weak Electrolytes. dose) close to the maximum solubilization capacity of
Weak electrolytes (such as ionic buffers) and strong the formulation (i.e., the red curve, Fig. 30). The po-
electrolytes (e.g., sodium chloride that may be included tential for precipitation is reduced when the CMC is
in a diluent) increase the polarity of aqueous solutions low and is therefore less prevalent for nonionic sur-
and, therefore, lead to the opposite effect on surfactant factants compared with ionic surfactants. Indeed, mi-
properties compared with cosolvents (i.e., electrolytes cellar formulations are much less susceptible to drug
decrease surfactant solubility and CMC, increase mi- precipitation on dilution than, for example, cosolvent
celle size, and typically increase surfactant solubiliza- systems, since solubility increases linearly with in-
tion capacity) (Attwood and Florence, 1983; Yalkowsky, creasing surfactant concentration above the CMC and
1999; Tadros, 2005). Equation 38 provides a means of therefore decreases linearly with linear sample
predicting the CMC of a surfactant in the presence of
a strong electrolyte:

LogCMC 5 LogCMCno salt 2 ks ms ð38Þ

where ks and ms are the salt constant (describing


salting-out potential; see forthcoming text) and the salt
molality, respectively (Ray and Nemethy, 1971; Carale
et al., 1994). The capacity of a salt to lower the
CMC correlates with its Hofmeister-like properties
(Hofmeister, 1888), where salts showing the greatest
ability to restructure water (multivalent ions and
those with small ionic radii and high charge densities)
exhibit the greater capacity to “salt-out” a surfactant
from solution.
The presence of charged ions in solution can also
shield electrostatic repulsive forces between charged Fig. 30. The effect of dilution of a surfactant solution containing a high or
ionic surfactants; hence, the effect of salts on the CMC a low drug concentration. Points on the dilution curves at which
theoretical drug concentrations exceed maximum solubility (labeled X
is comparatively greater for ionic surfactants compared and Y) indicate areas of precipitation risk. S0 is the intrinsic drug
with nonionic surfactants (Carale et al., 1994). The solubility in the absence of surfactant.
384 Williams et al.

dilution. In contrast, the exponential increase in TABLE 20


Selected oral and parenteral toxicity of some commonly used
solubilization capacity observed on addition of cosol- nonionic surfactants
vent results in nonlinear decreases in solubilization
Surfactant Median Lethal Dose (LD50) values
capacity on dilution and greater potential for
Brij 35 Mouse: 0.16 g/kg i.p
precipitation. Mouse: 1 g/kg i.v.
Surfactants are commonly added to both cosolvent Rat: 0.027 g/kg i.v.
and buffered formulations to reduce the risk of pre- Rat: 8.6 g/kg p.o.
Tween 20 Mouse: 1.42 g/kg i.v.
cipitation, regardless of an effect on intrinsic solubility Mouse: 37 g/kg p.o.
(Nassar et al., 2004; Dong et al., 2008; Sanghvi et al., Tween 40 Rat: 1.58 g/kg i.v.
Tween 60 Rat: 1.22 g/kg i.v.
2009; Tonsberg et al., 2011). For example, low HLB Tween 80 Mouse: 7.6 g /kg i.p.
block copolymers have shown recent utility in reducing Mouse: 4.5 g/kg i.v.
Rat:8 g/kg i.v.
drug precipitation on dilution of micellar formulations Mouse: 25 g /kg p.o.
and pH-adjusted formulations of amiodarone (Elhasi Labrasol Rat: .22 ml/kg day p.o.
et al., 2007) and an unnamed poorly water-soluble drug Labrafil Rat: .2 g/kg day p.o.
Cremophor EL Dog: 0.64 g/kg i.v.
(Dai et al., 2008b). The mechanisms of precipitation Mouse: 6.5 g/kg i.v.
inhibition may simply reflect the additional solubiliz- Rat: .6.4 g/kg p.o.
Cremophor RH40 Mouse: 12.5 g/kg i.p.
ing power gained from the presence of surfactant, even Mouse: .12.0 g/kg i.v.
at high dilution, since the CMCs of the block copoly- Rat: .16 g/kg p.o.
mers are low compared with traditional nonionic sur- Cremophor RH 60 Mouse: .12.5 g/kg i.p.
Rat: .16 g/kg p.o.
factants (the CMC of Pluronic F127 is 2.8  1026 M, Poloxamer 188 Mouse: 1 g/kg i.v.
whereas the CMC of vitamin E TPGS 1000 succinate is Rat: 7.5 g/kg i.v.
1.3  1025 M (Kataoka et al., 2001; Kabanov et al., Rat: 9.4 g/kg p.o.

2002; Adams et al., 2003). Alternatively, the block Data taken from Rowe et al. (2009).

copolymers may inhibit drug crystallization directly


(Dai et al., 2008a) in a manner analogous to the (Krzyzaniak and Yalkowsky, 1998). The typical exper-
inhibitory effects of nonionic high-molecular-weight imental design used was described in Section VI.F.2,
polymers on drug crystallization from supersaturated and studies of this type have revealed that increasing
solutions (Warren et al., 2010). the chain length of nonionic surfactants appears to
promote surfactant toxicity against red blood cells
F. Potential Pharmacological Properties of (Ernst and Arditti, 1980), although this toxicity in-
Nonionic Surfactants crease is not maintained above a chain length of
A relatively narrow range of nonionic surfactants approximately 14 carbons (Schott, 1973; Ferguson
has been used in parenteral drug delivery, a situation and Prottey, 1976). The nature of the hydrophobic
perpetuated by concerns surrounding toxicity and portion of a surfactant may also impact on cell pene-
pharmacological activity and the limited number of tration, and in this regard, triglycerides are bulkier
surfactants with preexisting regulatory approval. The than monoalkyl materials and, therefore, may reduce
adverse effects reportedly resulting from surfactant surfactant cell interaction. Nonionic surfactants of
administration are described in the following section, equivalent/similar hydrophobic character have also
noting that these effects may reflect surfactant com- been shown to exhibit wide differences in lytic effects,
position directly or the presence of lower quantities of implicating a role for the hydrophilic component in
reaction products from the surfactant synthesis in the dictating toxicity differences (Zaslavsky et al., 1978;
final product. Tragner and Csordas, 1987). For example, nonionic
1. Toxicity. Table 20 lists the LD50 values for a surfactants comprising hydrophilic segments made up
variety of nonionic surfactants commonly used for of polyoxyethylene (e.g., Brij and Solutol HS-15)
parenteral drug delivery (Rowe et al., 2009). The exhibit significantly less cellular toxicity compared
mechanisms by which surfactants mediate toxicity with a range of sugar-based surfactants (Shalel et al.,
are not well described; however, injury to the local 2002; Soderlind et al., 2003), and the toxicity toward
vasculature, blood cells, and other tissues may be an epithelial cell line of a series of commonly used
mediated through the capacity to adsorb to cells and to surfactants with similar C18 chain-length lipophilic
1) disrupt the integrity of the cell membrane (Shalel regions and only small differences in surface activity
et al., 2002) or 2) at higher concentrations (above the (CMC values were similar) varied more than 100-fold
CMC) to solubilize lipid membranes (Jones, 1999). The in the following rank order: Simulsol 4000 (least toxic)
potential for surfactant (or indeed any excipient) to , Cremophor EL , Tween 80 , Tween 20 , Solutol
promote toxicity associated with blood incompatibility HS-15 , TPGS 1500 (Maupas et al., 2011). It was
is commonly screened in vitro by monitoring hemolysis further proposed that this rank order reflected varia-
in human or animal red blood cells after incubation with tion in the size of the hydrophilic group, with bulkier
excipient (e.g., surfactant) or a complete formulation polyoxyethylene chains providing a greater steric
Strategies for Low Drug Solubility 385

barrier to cell membrane interactions (Maupas et al., to the hypersensitivity reactions associated with
2011). Cremophor compared with rodent animal models
Several nonionic surfactants also can cause more (Lorenz et al., 1982; Park et al., 2009). The i.v. use of
widespread systemic effects after parenteral use, and Cremophor-containing formulations has also been
the risks associated with the systemic administration associated with peripheral neuropathy (Wiernik
of Cremophor and Tween surfactants have been well et al., 1987; van Zuylen et al., 2001).
documented (Gelderblom et al., 2001; ten Tije et al., The oral LD50 for Cremophor EL and RH40 in rats is
2003). Indeed, the Taxol formulation (paclitaxel; 6.4 and 16 g/kg, respectively (Table 20) (Rowe et al.,
Bristol-Myers Squibb) has been the subject of consider- 2009). Rats dosed with 100 mg/kg Cremophor EL for 1
able discussion owing to its high Cremophor EL month showed no evidence of adverse reactions (Gad
content (see Table 19). Used primarily for the treat- et al., 2006), and in human subjects dosed with 5 g of
ment of breast cancer, Taxol is a nonaqueous solution Cremophor EL, only one in eight reported symptoms
that contains 30 mg of paclitaxel dissolved in 5 ml of that were attributed to the surfactant (in this case,
a mixture of 51% Cremophor EL and 49% dehydrated mild GI disturbance was observed) (Martin-Facklam
alcohol. It is diluted at least 20-fold with saline or et al., 2002b). Indeed, evidence of the tolerability of
5% dextrose before administration by i.v. infusion oral Cremophor can be drawn from the clinical use of
(Bristol-Myers Squibb Company, 2007). Despite Norvir SEC (ritonavir) oral capsules for the treatment
dilution and slow rates of infusion, however, acute of HIV where daily doses of Cremophor EL of ;1.2 g
hypersensitivity reactions are commonly observed and are common (Strickley and Oliyai, 2007).
manifest as dyspnea, flushing, urticaria, hypotension, Docetaxel (Taxotere) (Table 19) is formulated as
and angioedema (Weiss et al., 1990; Rowinsky et al., a nonaqueous solution in 100% Tween 80; like Taxol,
1991; Gelderblom et al., 2001; Singla et al., 2002). it is routinely diluted before i.v. administration
Paclitaxel (i.e., aside of any vehicle effect) may also (Strickley, 2004). Although Taxotere has been associ-
cause hypersensitivity reactions (Essayan et al., 1996); ated with some incidences of acute hypersensitivity
however, an increasing body of evidence suggests that (Schrijvers et al., 1993) and peripheral neuropathy
the hypersensitivity reactions associated with Taxol (Hilkens et al., 1997a,b), the frequency of adverse effects
are, at least in part, related to the Cremophor is lower than that of the Cremophor-containing for-
component of the formulation. Data to support this mulations, suggesting a tolerability benefit for Tween
contention includes evidence of 1) histamine release 80 (Gelderblom et al., 2001; ten Tije et al., 2003).
after administration of ethoxylated oleic acid (a minor The potential toxicity of some surfactants has
component of Cremophor EL) to dogs (Lorenz et al., limited wider application except at low concentrations.
1982), 2) complement pathway activation at Cremo- To this end, a significant effort has been directed
phor EL plasma concentrations above 2 mL/ml and toward reengineering older micellar formulations for
a reduction in the incidence of hypersensitivity at parenteral use, and a number of Cremophor-free pacli-
lower infusion rates that minimize Cremophor plasma taxel formulations have been developed to circumvent
concentrations (van Zuylen et al., 2000), and 3) the issues surrounding the use of Taxol (van Zuylen
hypersensitivity reactions after the i.v. administration et al., 2001; Singla et al., 2002). One notable example is
of other Cremophor-containing formulations, such as Genexol-PM (Samyang Corporation, Seoul Korea),
Sandimmune (Volcheck and Van Dellen, 1998). In the where paclitaxel is solubilized by micelles of a polylactic
latter study, the same patients showed no evidence of acid-polyethylene oxide block copolymer (Kim et al.,
allergic reaction on receiving oral doses of the same 2004b). Compared with Taxol, the maximum tolerable
formulations, providing additional support to the dose of Genexol in mice is three times higher (Kim
suggestion that Cremophor is either digested or not et al., 2001). The formulation was also able to solubilize
absorbed (or both) after oral administration and, the drug in a micellar form at up to 10 mg/ml, which is
therefore, that the toxicity concerns that arise after ;1.7-fold higher than the original Taxol formulation
parenteral administration are minimized (Volcheck (Kwon and Forrest, 2006; Zhang and Feng, 2006). Abra-
and Van Dellen, 1998; Gelderblom et al., 2001). xane (Celgene, Summit, NJ) is another Cremophor-free
Sensitivity to parenteral Cremophor now dictates that formulation for paclitaxel where the drug is bound to
patients receiving Taxol be premedicated with anti- albumin particles (Green et al., 2006). Abraxane is
histamines and corticosteroids to reduce histamine- currently approved by the FDA for the treatment of
related adverse reactions (Weiss et al., 1990; Pazdur breast cancer. A recent advance in the use of sur-
et al., 1993), although minor reactions (facial flushing factants for parenteral solubilization is the move to
and rash) remain common (Gelderblom et al., 2001). toward purified surfactants. For example, ultrapurified
Significant interspecies differences toward hypersensi- polysorbate 80 (HX2) is now available from the NOF
tivity reactions to Cremophor EL are also apparent. Corporation (White Plains, NY) for oral and parenteral
Dogs are highly sensitive to drugs and excipients that use. HX2 is derived from vegetable-based components
cause histamine release and are therefore more prone and is designed to offer a reduction in hypersensitivity
386 Williams et al.

and better stability compared with traditional poly- evidence of higher plasma concentrations in compari-
sorbate surfactants (Tween). son with the other formulations, suggesting that
Solutol HS-15 (BASF Corporation) has also been changes to drug clearance are possible, although in
investigated for use in oral and parenteral drug this case, changes were specific to Cremophor EL
delivery (Ku and Velagaleti, 2010). Solutol consists of rather than Tween 80 (Gelderblom et al., 2001). van
ethoxylated (15 moles) glycerides of 12-hydroxystearic Zuylen et al. (2001) have shown a similar increase in
acid and is much like the Cremophor-class of surfac- paclitaxel concentrations after coadministration with
tants. Similar levels of histamine release have been Cremophor EL and suggested that the mechanism
observed after administration of Solutol and Cremo- of enhanced exposure reflected the low volume of
phor to various animals (Ku and Velagaleti, 2010). distribution of Cremophor EL (;blood volume) and
Solutol has approval for use in injectable formulations subsequent entrapment of paclitaxel in surfactant
in Mexico, Argentina, and Canada (Strickley, 2004), micelles in the plasma (van Zuylen et al., 2001). It is
but in Europe and the US, it is still regarded as a novel not clear why this effect was not observed with Tween
excipient (Ku and Velagaleti, 2010). 80 (Fig. 31), although Tween 80 is cleared more quickly
2. Pharmacokinetic Effects. Changes in the phar- than Cremophor (ten Tije et al., 2003) and Cremophor
macokinetics of several compounds—including pacli- EL has been suggested to impact on excretory pro-
taxel, cyclosporine, doxorubicin, and etoposide—have cesses such as P-glycoprotein (P-gp) efflux (see Section
all been linked with the use of Cremophor in paren- XI). The existence of Cremophor micelles in the blood
teral formulations (Ellis et al., 1996; Millward et al., after administration of Taxol may also lead to altered
1998; van Zuylen et al., 2001; Bittner and Mountfield, pharmacokinetics for simultaneously administered
2002; Buggins et al., 2007), and Tween 80 has been drugs (ten Tije et al., 2003) and, therefore, the
shown to alter drug disposition after coadministration potential for drug interactions. Further descriptions
with doxorubicin, etoposide, and digoxin (Colombo of the effect of Cremophor EL on the pharmacokinetics
et al., 1996, 1999; Zhang et al., 2003). The general of paclitaxel and other drugs are found in the following
consensus to emerge from these studies is that the references: Gelderblom et al. (2001), ten Tije et al.
presence of nonionic surfactants (at high concentra- (2003), and Buggins et al. (2007).
tion) may lead to an increase in systemic exposure of Beyond simple physicochemical mechanisms of sol-
a coadministered drug compared with administration ubilization (and therefore changes to pharmacokinetics
in, for example, a cosolvent formulation (van Zuylen presumably via differences in free concentration), it
et al., 2001; ten Tije et al., 2003). By way of illustration, is increasingly apparent that surfactants (including
Fig. 31 shows the pharmacokinetics of paclitaxel dosed Cremophor) can affect the activity of metabolic cyto-
to mice in formulations containing Cremophor EL, chrome P450 enzymes and excretory drug transport-
ers, thereby potentially changing clearance and
Tween 80, or DMA (a cosolvent). Administration of
distribution processes directly (Bittner et al., 2003;
paclitaxel in the Cremophor EL formulation shows
Wandel et al., 2003; Bravo González et al., 2004;
Christiansen et al., 2011). For example, several drugs,
including many cytotoxic agents, are substrates for
efflux transporters such as P-gp, breast cancer resist-
ance protein (BCRP), and multidrug resistance pro-
teins (MRP1 and MRP2) (Schinkel et al., 1996; Borst
and Elferink, 2002; Leonard et al., 2003). Upregulation
of these transporters is evident in many tumor types,
leading to the phenomenon of multidrug resistance.
P-gp expression at the luminal side of the blood-brain
barrier (BBB) is also believed to limit drug access to
the brain and, therefore, the cytotoxity of some anti-
cancer drugs that are P-gp substrates (van Asperen
et al., 1996; Kemper et al., 2003). Cremophor EL has
been reported to inhibit P-gp and may therefore
reverse the implications of this MDR phenotype
(Friche et al., 1990b; Schuurhuis et al., 1990; Woodcock
et al., 1990). Similar P-gp inhibition has been shown in
several cancer cell lines using Tween 80 (Friche et al.,
1990b), Solutol HS-15 (Coon et al., 1991), TPGS
Fig. 31. Effect of the formulation vehicle on paclitaxel concentration in (Dintaman and Silverman, 1999; Bogman et al.,
female mice after an i.v. bolus administration of the drug at a concentra-
tion of 10 mg/kg. DMA, dimethylacetamide. Values are expressed as 2003), and Pluronic surfactants (Yang et al., 2007).
means 6 SEM. Adapted from Sparreboom et al. (1999). Cremophor EL, Tween 20, Span 20, Brij 30, and
Strategies for Low Drug Solubility 387

Pluronic P85 have all been shown in vitro to inhibit the in conjunction with solubilization approaches, includ-
efflux properties of BCRP (Kabanov et al., 2003; ing cosolvency, pH adjustment, and lipid-based de-
Yamagata et al., 2007a, 2009), and recent in vitro livery systems. However, realizing the potential for
studies have shown that Cremophor EL is able to surfactant-induced hypersensitivity after parenteral
inhibit MRP2 to a greater extent than probenecid (a administration (e.g., Cremophor EL in commercial
small molecular inhibitor of MRP2). In contrast, in the paclitaxel formulations) and the possibility of surfac-
same study, TPGS and Tween 80 showed only tant effects on free concentration and transporter/
moderate effects on MRP2 inhibition (Hanke et al., metabolic enzyme activity, care must be taken to
2010). interpret correctly the pharmacokinetic and pharma-
Coadministration of drugs that are substrates for codynamic data obtained in the presence of surfactants
efflux transporters, such as P-gp, BCRP, and MRP, (especially after parenteral administration).
with surfactants such as Cremophor therefore has the
potential to inhibit P-gp activity and promote access to VIII. Cyclodextrins
tumor cells, to the brain, or, in the case of brain
tumors, to both. To this end, rapidly (#6 h) infused A. Rationale for the Use of Cyclodextrins in the
doses of Taxol to human subjects have been shown to Solubilization of Poorly Water-Soluble Drugs
lead to plasma concentrations of Cremophor EL that Cyclodextrins (CDs) are macrocyclic oligosaccharides
are in excess of those seen to inhibit P-gp activity in consisting of a hydrophilic outer exterior and a hydro-
vitro by more than 50% (Webster et al., 1993; Rischin phobic inner cavity (Fig. 32). Drug molecules are
et al., 1996). However, later in vivo studies have been able to form dynamic inclusion complexes within this
unable to translate the in vitro inhibitory effects of cavity, and the higher solubility of the drug-CD comp-
Cremophor EL on P-gp activity into improved efficacy lex relative to the solubility of drug alone increases
for adriamycin (a P-gp substrate) in mice bearing MDR apparent solubility, often over several orders of mag-
leukemia (Watanabe et al., 1996) or to increase pacli- nitude. As a result, there has been widespread interest
taxel uptake into the mouse brain (Kemper et al., in the use of CDs to address low solubility in both oral
2003). The lack of correlation between in vitro and in and parenteral drug products (Loftsson and Brewster,
vivo endpoints may reflect differences in in vivo and in 1996; Rajewski and Stella, 1996; Stella and Rajewski,
vitro P-gp expression or difficulties in accurately re- 1997; Uekama et al., 1998; Davis and Brewster, 2004;
producing the physical and biochemical barrier prop- Brewster and Loftsson, 2007; Loftsson and Duchêne,
erties of the cell membrane in vitro. 2007). Examples of marketed formulations that use CD
In addition to potential effects on drug efflux from complexation are listed in Table 21 and include oral
the brain and from cancer cells, surfactants have been solid dosage forms (e.g., Omebeta, Betafarm) and liquid
widely studied for their effects on P-gp in the small dosage forms for both oral (Sporanox Oral Solution;
intestine (where P-gp may limit drug absorption) (Rege Janssen Pharmaceuticals) and parenteral (e.g., Vfend,
et al., 2002; Bogman et al., 2003, 2005) and on cyto- Pfizer) routes of administration.
chrome P450 enzymes in the small intestine that Common CDs include those formed by six, seven, or
may contribute to the first-pass metabolism of orally eight saccharide monomers, and these are classified as
administered drugs (Bravo González et al., 2004; a, b, or g CD, respectively. The most widely used CDs
Christiansen et al., 2011). Since surfactants in oral in pharmaceutical preparations are based on b-CD
drug delivery are typically used in combination with (i.e., CDs containing seven sugar moieties) (Fig. 32).
other excipients such as lipids and cosolvents, their The broad utility of b-CD stems from the good
capacity to affect efflux and first-pass metabolic alignment between b-CD cavity size and the aromatic
liabilities are discussed alongside effects on solubiliza- scaffolds commonly encountered in poorly water-soluble
tion in the section of this review focused on lipid based drugs. Since the upper limit of the solubility of a drug-
formulations (see Section XI). CD complex is that of the CD alone, the relatively poor
water solubility of unmodified CDs limits their wide-
G. Summary spread utility. Modified CDs, such as hydroxypropyl-
Above the critical micelle concentration, surfactants b-CD (HP-b-CD) and sulfobutylether-b-CD (SBE-b-CD)
increase drug solubility via micellar solubilization. The (Table 22), have therefore been introduced, the solubil-
lipophilic micelle core provides a nonpolar reservoir ity of both being significantly higher than that of the
into which highly lipophilic compounds may partition, unmodified material and typically in excess of 0.5 g/ml
enhancing apparent aqueous solubility. In addition, (see Table 22). The higher solubility of chemically
poorly water-soluble drugs may also be solubilized modified CDs also reduces the nephrotoxicity previously
by other regions of the micelle, in particular the associated with poorly water-soluble CDs (Brewster
polyoxyethylene-rich mantle. A large number of non- et al., 1989; Thompson, 1997; Stella and He, 2008). This
ionic surfactants have been used in both oral and improved toxicological profile of the modified CDs has
parenteral preparations, and surfactants may be used enhanced interest in the use of CDs in parenteral
388 Williams et al.

Fig. 32. The (A) chemical structure and (B) geometric configuration of a b-cyclodextrin ring. Adapted from Uekama et al. (2004).

applications. CDs are generally not absorbed after oral with altered drug pharmacokinetics after i.v. adminis-
administration; therefore, the potential for systemic tration of a drug-CD complex (Charman et al., 2006),
toxicity is much lower than that after parenteral although this appears to be the exception rather than
administration. Recent interest in the use of CDs as the rule and only likely for drug molecules with CD
actives in of themselves (e.g., effects on cholesterol binding constants in excess of 1  105 M21 (Stella and
absorption and renal drug elimintation) raises a number He, 2008).
of potential complications with respect to excipient clas- The nature of the drug-CD complex dictates that
sification and the traditional assumption that excipients solubilization with CDs is molecularly specific and
are “inactive.” This is increasingly evident across driven by the relative “fit” of the drug molecule (or
a number of excipient classes, including, for example, a portion of the molecule) within the CD cavity. CD
surfactants where activity in transporter and metabolic approaches are therefore less generic than solubility
inhibition is increasingly described. However, the enhancement strategies such as micellar solubilization.
ramifications of this with respect to drug product Nonetheless, CDs solubilize a wide variety of structur-
registration are yet to unravel fully. ally distinct drug molecules, both neutral and ionized,
The utility of CDs in assisting the delivery of poorly and may be used with other solubilization strategies,
water-soluble drugs is dictated by the affinity of the such as pH adjustment.
drug for the CD as indicated by the magnitude of Solid drug-CD complexes may also be isolated by, for
the drug-CD binding constant. Increasing affinity in- example, freeze-drying drug-CD solutions. Parenteral
creases the proportion of the total solubilized drug drug-CD formulations that can be reconstituted before
concentration that is present within the CD complex use and oral formulations containing solid CD-drug
and reduces the quantity of CD required to dissolve a complexes are therefore possible. In such formulations,
target drug mass. Increased drug affinity for the rates of in vitro dissolution of the drug-CD complex
complex is therefore favored in most circumstances. may be dramatically increased compared with uncom-
High binding affinities, however, may lead to incom- plexed drug due to the significant increases in
plete dissociation of the drug-CD complex after apparent solubility of the drug-CD complex. The
administration (and dilution) and limit the concentra- generation of an altered solid form of the drug also
tion of free drug available for absorption or activity. provides additional mechanisms of dissolution en-
Very high binding affinities have also been associated hancement via changes to solid-state properties.

TABLE 21
Commercially available cyclodextrin-containing formulations
Drug/Product Cyclodextrin (CD) Used Route of Administration

Sporanox (itraconazole; Janssen Pharmaceuticals) HP-b-CD Oral solution/i.v. solution


MitoExtra, Mitozytrex (mitomycin; Novartis Pharmaceuticals) HP-b-CD i.v. infusion
Abilify (aripiprazole; Bristol-Myers Squibb and Otsuka Pharm) SBE-b-CD i.m. solution
Vfend (voriconazole; Pfizer) SBE-b-CD i.v. solution
Yaz (ethinylestradiol and drospirenone; Bayer) b-CD Tablet
Omebeta (omeprazole; Betafarm) b-CD Tablet
Caverject Dual (alprostadil; Pfizer) a-CD i.v. solution
Nitropen (nitroglycerin; Nippon Kayaku) b-CD Sublingual tablet
HP, hydroxypropyl; SBE, sulfobutylether.
Strategies for Low Drug Solubility 389
TABLE 22
Examples of commonly used cyclodextrins to aid oral and parenteral delivery of poorly water-soluble drugs

Cyclodextrin Common Name MSa and Mwb Aqueous Solubility


(mg/ml) (25°C)

a-CD (6 glucose units) Alfadex 2/972 14.5


b-CD (7 glucose units) Betadex 2/1135 1.85
g-CD (8 glucose units) Gammadex 2/1297 23.2
2-Hydroxypropyl-a-CD (HP-a-CD) Cavasol W6 HP (Wacker, Germany) 0.65/1199 .500
2-Hydroxypropyl-b-CD (HP-b-CD)c Cavasol W7 HP (Wacker, Germany) 0.65/1400 .600
Sulfobutylether-b-CD (SBE-b-CD)c Capitsol (CyDex Pharmaceuticals, USAd) 0.9/2163 .600
Randomly methylated-b-CD (RM-b-CD) Cavasol W7 M (Wacker, Germany) 1.8/1312 .600
Maltosyl-b-CD (Mal-b-CD) Ensuiko Sugar Refining Co. (Japan) 0.14/1459 .500
CD, cyclodextrin; HP, hydroxypropyl; SBE, sulfobutylether.
a
Average molar substitution.
b
Average molecular weight.
c
Also described as “hydroxypropyl betadex” and “betadex sulfobutyl ether sodium” by the United States Pharmacopeia and National Formulary.
d
CyDex Pharmaceuticals was recently acquired by Ligand Technology, La Jolla, CA.

Indeed, CD formulations where drug is present in the CD cavity. Changes in entropy (DS) are usually mod-
amorphous form have been described, and under these erate and either negative or slightly positive Komiyama
circumstances, amorphous CD formulations have much and Bender, 1978; Harrison and Eftink, 1982; Inoue
in common with, and share the potential advantages et al., 1993). In contrast, the association of two non-
of, solubilizing solid dispersion formulations (see polar molecules in solution (or, for example, amphi-
Section X) and microporous adsorbent formulations philes during micelle formation) is more commonly
(see Section XII.A). However, the disadvantages of associated with an increase in entropy as structured
formulations that contain drug in the noncrystalline water molecules are redistributed into free solution
form also apply to amorphous CD formulations; the (Frank and Evans, 1945). The more moderate changes
risks of drug precipitation post dissolution and in entropy during drug-CD complexation have been
amorphous-crystalline solid-state transitions during suggested to reflect a loss of rotational freedom on
storage are ever present. complex formation that offsets the gain in entropy
In this section, the fundamental principles under- during water redistribution. The free energy of com-
pinning the use of CDs to increase drug solubility are plexation is therefore driven by the negative enthalpy
reviewed and the drug and non-drug related factors term rather than by an increase in entropy and pro-
that affect drug-CD complexation discussed. Examples vides an example of enthalpy-entropy compensation
of CD utility in oral and parenteral delivery are sub- (Inoue et al., 1993; Rekharsky and Inoue, 2000; Omar
sequently presented, including a brief discussion of the et al., 2007; di Cagno et al., 2011). Ionic substituents
issues surrounding CD toxicity. Finally, a review of the located on the exterior segment of some CDs, for
mechanisms of drug release from the CD complex is example, SBE-b-CD, may also interact with the drug
presented and a brief assessment provided of the via ion-dipole interactions or via an ion pair if the
most common approaches to the manufacture CD drug is oppositely charged. Drug interactions external
formulations. to the cavity may also play a role in dictating the
stoichiometry of drug-CD complexation (Okimoto
B. Drug-Cyclodextrin Inclusion Complexes et al., 1996).
1. Binding Equilibria between a Drug and Cyclodextrin. Drug binding affinity to a CD is described by the
Complexation between a drug molecule (the “guest”) equilibrium binding constant (K), also referred to as
and CD molecule (the “host”) initially requires rear- the stability constant. The binding constant provides
rangement or displacement of water molecules that are a measure of the solubilizing potential of a particular
present within the CD cavity and those that solvate the CD toward the drug of interest. For a 1:1 complex of
drug molecule in solution. The complex is then formed drug-CD, the binding constant K is defined by eq. 39,
via intermolecular (noncovalent) interactions between where [Df] and [CDf] are the concentration of the free
the guest (drug) and the CD, and water molecules sub- drug and CD, respectively, and [D:CD] is the concen-
sequently reorganize around the newly formed drug- tration of the formed complex:
CD complex (Bekers et al., 1991).
The process of drug-CD complexation is typically K1:1 ½D : CD
characterized by a large and negative enthalpy (DH) Df 1 CDf ⇌ D : CD K 1:1 5    ð39Þ
Df CDf
(Ross and Rekharsky, 1996; Rekharsky and Inoue,
2000; Liu and Guo, 2002) that reflects the intermolec- From eq. 39, it is clear that the higher the binding
ular interactions that favor complex formation. These constant (K), the greater the ratio of complexed drug to
interactions usually comprise van der Waals and hydro- uncomplexed drug. Binding constants typically lie
phobic interactions between the drug and the hydrophobic between 0 (no binding) and 1  106 M21 (high binding),
390 Williams et al.

although most drugs fall within the range of 1  102 are determined in solutions of increasing CD concen-
and 1  104 M21 (Rajewski and Stella, 1996; Stella and tration. An example phase-solubility profile is shown in
Rajewski, 1997; Stella et al., 1999). Higher binding Fig. 33A, with the gradient of the slope reflecting the
constants are desirable since lower levels of CD may be solubilizing capacity of the CD. The binding constant
used in the formulation; however, there is a risk of for a 1:1 complex is calculated by solving eq. 42 using
incomplete dissociation of the drug-CD complex after the linear phase solubility profile (gradient m, in-
administration if the binding constant is very high tercept S0):
(i.e., in excess of 1  105 2 1  106 M21). Incomplete
dissociation may in turn affect drug pharmacokinetic K 1:1 S0
ST 5 S0 1 ½CDT  5 m½CDT  1 S0 ð42Þ
and pharmacodynamic properties (Stella and He, 1 1 K 1:1 S0
2008). This is described in more detail in Section
Rearranging to generate eq. 43:
VIII.E.
In saturated solutions (i.e., at equilibrium), [Df] in K 1:1 S0
eq. 39 is the intrinsic drug solubility (S0). The total m5 ð43Þ
1 1 K 1:1 S0
concentration of dissolved drug in a CD solution (ST) is
therefore the sum of free drug concentration and drug Eq. 43 can itself be arranged to calculate the binding
complexed with the CD: constant, K:
m
ST 5 S0 1 ½D : CD ð40Þ K 1:1 5 ð44Þ
S0 2 m
Merging eq. 39 and eq. 40 provides an expression
that indicates the solubility gain that may be achieved Other commonly used methods to determine drug-
at different CD concentrations: CD binding constants include NMR spectroscopy, UV
spectroscopy, circular dichroism, potentiometry, micro-
K 1:1 S0 calorimetry, freezing-point depression, HPLC/TLC,
ST 5 S0 1 ½CDT  ð41Þ membrane permeation techniques, and capillary elec-
1 1 K 1:1 S0
trophoresis. Spectroscopic techniques (e.g., NMR) have
where [CDT] is the total CD concentration added to the proved useful in discerning drug-CD complex stoichi-
aqueous solution. CD concentrations in pharmaceuti- ometry (Bettinetti et al., 1991; Piel et al., 1998; Miyake
cal preparations greater than 40% are unusual (Li and et al., 1999b; Ribeiro et al., 2005a; di Cagno et al.,
Zhao, 2007; Stella and He, 2008). Therefore, if the 2011). Provided the environmental factors likely to
intrinsic drug solubility (S0) is known, eq. 41 allows affect solubility (e.g., pH and temperature) are tightly
rapid assessment of the potential utility of CDs controlled, different methods may generate highly
through substitution of the target drug concentration consistent results (Kurkov et al., 2011). However, in
for ST and calculation of the required value for K to some cases, higher-order complexes identified during
achieve this target drug concentration. Where the phase solubility studies have not been able to be veri-
value for K exceeds values typical for pharmaceutical fied by NMR experiments or UV spectroscopy. These
molecules (i.e., . 1  106 M21), CD alone is unlikely to discrepancies may be attributed to the more dilute
provide a suitable solubilization strategy. In these conditions used in the NMR/UV spectroscopy studies,
situations, an alternative strategy or a combined ap- in which there is a reduced likelihood of forming
proach (e.g., CDs plus pH adjustment in the case of an higher-order complexes in comparison with the higher
ionizable drug) may be required. It should be noted concentration conditions in phase-solubility experi-
that eq. 41 applies only for drugs that form 1:1 ments (Loftsson et al., 2002; Magnusdottir et al.,
complexes. The formation of a higher-order (e.g., 2:1) 2002; Loftsson et al., 2005).
drug-CD complex requires less CD to achieve the in- Phase-solubility profiles are classified according to
tended target concentration. Similarly, lower order their shape (Higuchi and Connons, 1964), and vary
complexes (e.g., 1:2 drug-CD complexes) suggest poor depending on the solubility and stoichiometry of the
drug CD complexation and the need for larger complex (Fig. 33). Two main solubility profiles are
quantities of CD to facilitate effective solubilization. evident and are described as either A-type or B-type.
Where the binding constant of a drug of interest has B-type profiles typically lead to limited increases in
already been determined or is available in the drug solubility and are more commonly (but not ex-
literature, the CD concentration required to dissolve clusively) associated with nonsubstituted CDs (Szejtli,
the drug may also be calculated from eq. 41. 1994; Brewster and Loftsson, 2007). BS-type profiles are
2. Measurements of Drug-Cyclodextrin Binding reflective of situations where complexation initially re-
Constants. The binding constant of a drug-CD com- sults in an increase in drug solubility, but the maximum
plex can be measured using several different appro- solubility of the complex in solution is soon reached (the
aches. In one of the most common approaches, the plateau phase in Fig. 33B). Further increases in CD
phase solubility approach, dissolved drug concentrations concentration result in precipitation of the complex,
Strategies for Low Drug Solubility 391

Fig. 33. Phase-solubility profiles with increasing cyclodextrin (CD)


concentration. (A) An example of a linear phase-solubility profile and Fig. 34. The effect of drug:cyclodextrin (CD) complexation ratio on the
(B) various types of phase-solubility profiles (described in the text). S0 is shape of AL and AP-type solubilization profiles. AL-type profiles are linear
the intrinsic drug solubility in the absence of CD. across the CD concentration range. Slopes of unity or less than unity
suggest the formation of 1:1 drug:CD complexes. Slopes of .1 suggest
formation of higher-order complexes with respect to the drug (e.g., 2:1
drug:CD). AP-type profiles suggest the formation of higher order
decreasing solubilized drug concentrations (Szejtli, complexes with respect to the CD (e.g., 1:2 or 1:3 drug:CD) at higher
CD concentrations. Additional techniques are required to verify that
1994; Brewster and Loftsson, 2007). As seen in Fig. solubility increases are due to drug:CD complexation and the stoichiom-
33B, BI-type profiles are characteristic of situations etry of the formed complex. S0 is the intrinsic drug solubility in the
absence of CD.
where the complex has lower solubility than drug alone
(Uekama et al., 1998; Brewster and Loftsson, 2007).
The lack of solubility gain makes B-type CD complexes
less attractive for most applications. 34). Slopes of gradient unity indicate 1:1 drug-CD
In contrast, significant increases in the quantity of complexes. Slopes that are greater than unity suggest
drug dissolved (ST) with increasing CD concentration the formation of a higher-order complex with respect to
are evident in A-type phase solubility profiles. A-type the drug (i.e., more than one drug molecule per CD
profiles that increase linearly with increase in CD molecule) (Yalkowsky, 1999; Brewster and Loftsson,
concentration are common within pharmaceutical 2007; Loftsson and Brewster, 2010).
sciences and are denoted AL-type. Examples of poorly Nonlinear increases in ST above a critical CD
water-soluble drugs exhibiting AL-type CD solubility concentration are also possible and are classified as
profiles include nicardipine hydrochloride with b-CD AP-type (i.e., a positive deviation) or AN-type profiles
(Fernandes et al., 2002), itraconazole with b-CD (Al- (i.e., a negative deviation) (Higuchi and Connons, 1964)
Marzouqi et al., 2006), indomethacin with b-CD (Fig. 33B). AP-type profiles are less common than AL-
(Jambhekar et al., 2004), cinnarizine with HP-b-CD type profiles but have been observed for several drugs,
and SBE-b-CD (Jarvinen et al., 1995), phenytoin with including cyclosporine A (Miyake et al., 1999a), itra-
HP-b-CD and SBE-b-CD (Savolainen et al., 1998), conazole (Miyake et al., 1999b; Peeters et al., 2002;
danazol with HP-b-CD (Badawy et al., 1996), dexa- Brewster et al., 2008a), miconazole (Pedersen et al.,
methasone with HP-b-CD (Loftsson et al., 1994), and 1993; Okimoto et al., 1996; Ribeiro et al., 2008), and
flavopiridol with HP-b-CD (Li et al., 1998a). tacrolimus (Arima et al., 2001b). AP-type profiles re-
The slope of AL-type profiles can provide an in- veal greater increases in drug solubilization than
dication of the stoichiometry of complex formation (Fig. expected from simple 1:1 complexes and therefore
392 Williams et al.

suggest the presence of higher-order complexes with (S0) of poorly water-soluble drugs is low and may be
respect to the CD (on reaching a certain CD concen- overestimated by the intercept in a phase solubility
tration) (Fig. 34). The stoichiometry of higher-order curve (Loftsson et al., 2005, 2007a; Loftsson and
complexes is inferred from the extent of the deviation Brewster, 2010).
in AP-type profiles (Brewster and Loftsson, 2007) and 3. Secondary Equilibria: Micelles and Cyclodextrin
the goodness of fit to either quadratic (for 1:2 com- Aggregate Formation. All natural cyclodextrins self-
plexes, eq. 45) or cubic (for 1:3 complexes, eq. 46) func- associate in solution to form aggregates and micelles
tions (Peeters et al., 2002): (Mele et al., 1998; Szente and Szejtli, 1999; Loftsson
et al., 2002; Bonini et al., 2006). Chemically modified
ST 5 S0 1 K 1:1 S0 ½CD 1 K 1:1 K 1:2 S0 ½CD2 ð45Þ CDs (e.g., HP-b-CD and SBE-b-CD) may also acquire
surfactant-like properties after incorporation of a hy-
ST 5 S0 1 K 1:1 S0 ½CD 1 K 1:1 K 1:2 S0 ½CD2 drophobic/lipophilic drug molecule (Loftsson et al.,
ð46Þ
1 K 1:1 K 1:2 K 1:3 S0 ½CD3 2002; Magnusdottir et al., 2002; Kurkov et al., 2011)
or, in the case of CDs with hydrophobic modifications,
The accuracy of these models is limited by the in the absence of a guest (Auzely-Velty et al., 2000). In
assumption that the total CD concentration is equal to addition to forming typical guest-host molecular inclu-
the free (uncomplexed) CD concentration. Errors occur sion complexes, CDs also have the potential to form
where a significant proportion of the CD is in the micellar aggregates and provide additional sites and
complexed form (as is often the case) (Peeters et al., mechanisms of drug solubilization.
2002; Brewster and Loftsson, 2007; Brewster et al., For example, phase solubility profiles of the sodium
2008a). Aside from higher-order molecular complexes salts of the nonsteroidal anti-inflammatory drugs
with respect to the CD, AP-type profiles may also result ibuprofen and diflunisal with HP-b-CD have been
from the formation of noninclusion complexes or via reported to be AL type with slopes of greater than 1,
solubilization of drug by drug-CD aggregates. These indicative of the formation of higher-order (e.g., 2:1)
potential situations are described in more detail in drug-CD complexes (Magnusdottir et al., 2002; Kurkov
Section VIII.B.3. et al., 2011). Space-filling docking analysis of the 1:1
Phase solubility profiles showing negative devia- complex, however, suggested that the CD cavity in HP-
tions, or AN-type profiles (Fig. 33B), describe situations b-CD was too small to accommodate an additional drug
where the solubilization efficiency of the CD decreases molecule, and the authors concluded that HP-b-CD
on reaching a critical concentration (Higuchi and solubilization of ibuprofen and diflunisal was a com-
Connons, 1964). This may occur as a result of 1) a bination of molecular complexation and drug solu-
decrease in solvent quality through increased solution bilization via aggregates of the drug-CD complex
viscosity at higher drug and CD concentrations, 2) self- (Magnusdottir et al., 2002; Kurkov et al., 2011).
association of free CD (thereby reducing the available Similarly, the shape of phase solubility profiles for
concentration of free CD), or 3) where ionic interactions cholesterol–HP-b-CD and cholesterol–SBE-b-CD could
are evident, CD-mediated changes in the apparent pKa not be explained by the formation of drug-CD inclusion
of the drug (Uekama et al., 1998; Brewster and complexes alone (Loftsson et al., 2002), prompting the
Loftsson, 2007). AN-type profiles are comparatively suggestion of a role for drug solubilization via non-
rare in the pharmaceutical literature but have been inclusion complexation in the overall solubility profile.
observed for miconazole (Piel et al., 1998; Ribeiro et al., Further evidence that cholesterol-HP-b-CD complexes
2008), carvedilol (Miro et al., 2006), and gefitinib (Lee may have intrinsic solubilization properties is appar-
et al., 2009). ent in studies showing an increase in cyclosporine A
The complexation efficiency (CE) provides an alter- solubilization by HP-b-CD on the addition of cholesterol
native to the binding constant for quantification of CD (Loftsson et al., 2002, 2005). In the latter case, the
solubilization capacity (Loftsson et al., 2007a). CE is increase in cyclosporine A solubility occurred despite
calculated from the concentration ratio between com- the expectation that the addition of cholesterol would
plexed and uncomplexed CD: result in a decrease in drug solubilization (as a result of
competition for the CD cavity). A number of methods
½D : CD m have also been proposed to allow detection of CD
CE5 5S0 × K 1:1 5 ð47Þ aggregates directly (rather than via indirect inference
½CD ð1 2 mÞ
from phase solubility studies), and these methods have
A CE value of 1 indicates that all added drug is been the subject of a recent review (Messner et al.,
complexed with CD (in a 1:1 ratio), whereas a CE of 0.1 2010). It remains unclear, however, whether drug-CD
indicates that 1 drug molecule of 11 has formed a aggregates affect the in vivo performance of CD
complex (Loftsson et al., 2005). CE has been suggested formulations since the aggregates are held together
to provide a more robust indication of the solubilization through weak interactions such as hydrogen bonding
properties of different CDs since the intrinsic solubility and van der Waals interactions, and deaggregation on
Strategies for Low Drug Solubility 393

dilution in blood or in the GI fluids is likely (Messner As only a portion of the drug molecule is required to
et al., 2010, 2011b). form a drug-CD complex, and since many molecules
contain more than one aromatic group, larger drug
C. Drug Factors Affecting Complexation molecules are well suited to forming higher-order
by Cyclodextrins complexes with respect to the CD (i.e., one drug
1. Size. As different CDs show differences in the molecule and multiple CD molecules). Itraconazole
size of the hydrophobic cavity and the degree and can form 1:2 complexes with HP-b-CD, and under
nature of chemical substitution, drug affinity toward neutral conditions, even 1:3 complexes are possible
a range of CD molecules may vary considerably. (Miyake et al., 1999b; Peeters et al., 2002).
However, drug molecular size, relative to the size of One notable example of the use of CDs other than
the hydrophobic cavity, provides a strong indication of b-CD is the development of sugammadex (Bridion,
the likelihood of complexation and also the order of the formerly Organon, now Merck). Sugammadex is a mod-
complex (Loftsson et al., 1993; Loftsson and Brewster, ified g-CD with a very high binding affinity for the
1996; Mura et al., 1999; Loftsson et al., 2002, 2005; anesthetic rocuronium [the sugammadex-rocuronium
Tønnesen et al., 2002). binding constant is in the order of 1  107 M21 which is
The size of the CD cavity varies depending on the at least 10-fold higher than any of the other drug-CD
number of glycopyranose monomers in the ring. Thus, binding constants in the literature (Bom et al., 2002)].
a-CDs have the smallest cavity (;4.7–5.3 Å in The high binding affinity of sugammadex for rocuro-
diameter) and g-CDs the largest (;7.5–8.3 Å in nium ensures that i.v. administration of sugammadex
diameter). The size of the inclusion complex cavity results in rapid sequestration of rocuronium into the
for b-CD is ;6.0 – 6.5 Å (Uekama, 2004). CD complex and effective reversal of neuromuscular
The size of the hydrophobic cavity of the CD host blockade without the need for the administration of
determines the fit of the guest drug molecule (or anticholinesterases.
molecular fragment) and, subsequently, the extent of 2. Polarity and Charge. As the inner cavity of a CD
the interaction of the guest with the host post- molecule is hydrophobic, increasing drug hydrophobic-
insertion. The dependence of the size of the CD cavity ity typically increases binding affinity (Bergeron et al.,
on the generation of ideal guest-host complexes is well 1978; Uekama et al., 1982, 1983; Jones and Parr, 1987;
exemplified by Cromwell et al. (1985), who examined Marques et al., 1990; Liu and Guo, 2002). Solubility
the binding affinity of a carboxylate derived adaman- enhancement via CD solubilization is therefore more
tane for a-CD, b-CD, and g-CD. The carboxylate- pronounced for drugs with increasing hydrophobicity.
derived adamantane was able to form complexes with For example, the solubility of dithiolethione (log P 5
both b-CD and g-CD, but a better fit with b-CD led to 1.58) increases ;9-fold on the addition of 10% HP-
an increased number of van der Waals interactions b-CD, whereas the solubility of the structurally related
between the guest and host and, ultimately, a higher but more lipophilic 5-phenyldithiolethione (log P 5
enthalpy of binding. This in turn favored complexa- 3.67) and anetholetrithione (log P 5 3.82) increases
tion with the b-CD derivative, whereas the adaman- ;400-fold in the presence of HP-b-CD (Dollo et al.,
tane was too large to complex effectively with the 1999). A similar trend was observed in the same study
smaller a-CD hydrophobic cavity (Cromwell et al., using SBE-b-CD. Albers and Muller (1992) also noted
1985). an increase in the effectiveness of HP-b-CD complex-
In addition to the complexation of adamantanes (for ation for more hydrophobic derivatives of testosterone
which b-CD has particularly high affinity), the b-CD and estradiol. Structural changes that increase hydro-
cavity is well suited to the insertion of aromatic phobicity, such as increasing alkyl chain length or the
benzene rings. Since this functional group is found size and number of aromatic rings in the drug mol-
widely in the molecular structure of many drugs b-CD, ecule, are likely to contribute to enhanced CD com-
and its derivatives, are well suited for solubilization of plexation, although structural modification of this type
a large number of pharmaceutical molecules (Marques will also reduce intrinsic solubility (in water), thereby
et al., 1990; Nakai et al., 1990; Tong et al., 1991; increasing the need for enhanced solubilization. The
Kearney et al., 1992; Takisawa et al., 1993). Impor- net effect of increasing drug hydrophobicity on total
tantly, the entire drug molecule does not need to enter drug solubility is therefore difficult to predict.
the hydrophobic cavity for effective CD complexation For ionizable drugs, the degree of complexation of
since binding may be achieved in instances where only the un-ionized form is typically greater than the
a small portion of the drug molecule is associated with ionized (and more polar) equivalent (Tinwalla et al.,
the CD. This is evident in the CD complexation of high 1993; Okimoto et al., 1996; Redenti et al., 2000).
Mw compounds such as tacrolimus (Arima et al., 2001b) Binding constants can be measured as a function of pH
and itraconazole (Peeters et al., 2002), both of which to describe the relative binding affinities of the un-
showed good binding to b-CD but are too large to be ionized (Ku) and ionized (Ki) forms (Okimoto et al.,
entirely enclosed within the CD cavity. 1996). Comparison of the degree of CD complexation
394 Williams et al.

with the un-ionized/ionized drug fraction (i.e., Ku/Ki) et al., 2002; Ran et al., 2005; He et al., 2006). This
provides an indication of the loss of binding affinity situation is analogous to the combined use of pH mani-
through the introduction of charge on the drug pulation with cosolvents and surfactants, where in-
molecule. Scheme 3 illustrates the effect of ionization creasing ionization reduces the impact of the cosolvent
on complexation between a neutral CD and an acidic or or micellar solubilization effect, but the net effect is to
basic drug. Although both ionized and un-ionized forms enhance total solubility.
of a drug may form an inclusion complex with the CD, An example of this synergy is provided in Fig. 35,
it is more common that the more hydrophobic un- where the solubility of the weak base flavopiridol
ionized form more readily complexes— and thus, Ku . increases with increasing acidity (i.e., with decrease in
Ki. The scheme also highlights the potential for pH) and with increases in HP-b-CD concentration at
dissociation of drug in free solution (Ka) and as the both high and low pH, suggesting that both un-ionized
complex (Ka’). and ionized drugs show some affinity for the CD (Li
From Scheme 3, total drug solubility under con- et al., 1998a). The marketed Sporanox formulations
ditions of changing pH is the sum of four species (i.e., provide further evidence of the impact of combining
the total of free and complexed drug) in both the un- pH with CD complexation. The i.v. formulation contains
ionized form and the ionized form (Tinwalla et al., 10 mg/ml itraconazole complexed with 40% HP-b-CD,
1993; Okimoto et al., 1996): with the pH adjusted to pH 4.5, whereas the less
For a weak acid: stringent irritancy concerns associated with oral ad-
ministration allow the oral formulation to be pH
ST 5½HA 1 ½HA : CD 1 ½A 2  1 ½A 2: CD ð48Þ
adjusted to pH 2.0 (Strickley, 2004).
For a weak base: The potential for CD ionization adds additional
    complexity to the understanding of pH effects on
ST 5½B 1 ½B : CD 1 BH 1 1 BH 1: CD ð49Þ drug-CD complexation. SBE-b-CD, for example, is
anionic and has been suggested to form ion pairs with
Although a pH change resulting in increased drug
the weakly basic drugs prazosin, papaverine, and
ionization may reduce CD binding affinity, the higher
miconazole in pH conditions where both drug and CD
aqueous solubility of the ionized form is often sufficient
are (partially) ionized (Okimoto et al., 1996; Zia et al.,
to compensate for decreased CD association, and total
2001). In contrast, for weakly acidic drugs (e.g., napro-
solubilized drug concentrations increase (Li et al.,
xen, warfarin, and indomethacin), the situation was
1998a, 1999c; Piel et al., 1998; Jain et al., 2001; Ni
reversed, and binding affinity of the charged species
was reduced. The authors suggested that this may be
attributed to electrostatic repulsion between SBE-
b-CD and the similar charged anionic drugs (Okimoto
et al., 1996; Zia et al., 2001). Electrostatic repulsion
may therefore limits complexation, although increas-
ing the ionic strength of the solution may restore some
degree of association via salting out and charge

Scheme 3. Complexation of ionized and un-ionized forms of an acidic


drug (HA, top scheme) and a basic drug (B, lower scheme) to a neutral
cyclodextrin (CD) in a 1:1 complex. Ka/Ka’ is the dissociation constant
between ionized and un-ionized forms of the drug in the free uncomplexed Fig. 35. Total experimental aqueous flavopiridol solubilities in hydrox-
form/complexed form. Ku/Ki is the drug-CD binding constant for the ypropyl-b-cyclodextrin (HP-b-CD) solutions at different pH values.
unionized/ionized form of the drug. Adapted from Okimoto et al. (1996). Adapted from Li et al. (1998a).
Strategies for Low Drug Solubility 395

shielding effects (Zia et al., 2001; Brewster et al., D. Impact of Cosolubilizers and Excipients on
2008a). Cyclodextrin-Mediated Drug Solubilization
Finally, a change in solution pH may also influence 1. Cyclodextrins and Water-Soluble Polymers.
the order of drug-CD complexes. Itraconazole is a weak Many hydrophilic polymers increase drug binding to
base and complexes with HP-b-CD in a pH-dependent CDs. The addition of HPMC or PVP increases
manner, as the three potential proton acceptor sites vinpocetine binding to b-CD and SBE-b-CD (Fig. 36)
are, under conditions of increasing pH, progressively (Ribeiro et al., 2003b, 2005a), and HP-b-CD binding
deprotonated and, therefore, increasingly available for constants for hydrocortisone, dexamethasone, and
interaction with CDs. Increases in pH from 1.5 to 2.5 to carbamazepine are all increased in the presence of
4.0 leads to a progressive increase in the K1:2 binding HPMC (Loftsson et al., 1994; Smith et al., 2005; Kou
constant (i.e., an increase in affinity for the 1:2 drug- et al., 2011). Similarly, the solubility of D9-tetrahydro-
CD complex) (Miyake et al., 1999b; Peeters et al., cannabinol increases approximately 200-fold in the
2002), and molecular mechanics simulations suggest presence of 8% HP-b-CD and approximately 300-fold in
the formation of a 1:3 drug-CD complex at neutral pH solutions containing 8% HP-b-CD and 0.1% HPMC
(Peeters et al., 2002).
The potential for electrostatic repulsion is generally
thought to preclude charged CDs from forming higher-
order complexes (with respect to CD) (Thompson, 1997;
Loftsson et al., 2002); however, evidence of AP-type
increases in itraconazole solubility in the presence of
SBE-b-CD (Brewster et al., 2008a) suggest that this may
not always be the case. Whether AP-type solubilization
profiles using charged CDs are due to conventional
complexation or solubilization via noninclusion com-
plexes is currently not well understood and is likely to be
further complicated in situations where the drug itself
may be ionized (as ion-pair interactions between drug
and CD are likely to contribute to drug solubilization).
3. Presence of Counterions. Several studies have
reported an increase in CD solubilization for ionizable
compounds in the presence of low-molecular-weight
acidic/basic counterions, including tartaric and malic
acid, bile acids, and various amino acids (Piel et al.,
1997; Faucci et al., 2000; Mura et al., 2001b; Mura
et al., 2003, 2005). The mechanism of solubilization
enhancement is attributable to an ion-pair formation,
where the CD and counterion interact via a “salt
bridge” that stabilizes the drug-CD complex and pro-
motes solubilization of ionized drug.
For example, naproxen solubility in HP-b-CD is
higher in the presence of arginine (a basic counterion)
(29.5 mg/ml) compared with HP-b-CD solutions of
equivalent pH without arginine or in solutions contain-
ing arginine alone (11.3 mg/ml and 14.5 mg/ml, re-
spectively) (Mura et al., 2003). Solubility advantages are
also dependent on the nature of the counterion, and the
solubility of an econazole-malic acid-a-CD ternary
complex has been reported to be enhanced compared
with equivalent ternary complexes formed in the
presence of drug, CD, and lactate or citrate counterions
(Mura et al., 2001b). Since the lactate and citrate salts
of econazole were more soluble than the malate salt in
the absence of CD, the authors concluded that steric
Fig. 36. Phase-solubility diagrams for vinpocetine at room temperature
effects and counterion size (and not salt solubility) were in the presence of (A) b-cyclodextrin (b-CD) and (B) sulfobutylether-
more important factors in determining the extent of b-cyclodextrin (SBE-b-CD) with or without the water-soluble polymers
HPMC/PVP. Values are expressed as means. HPMC, hydroxypropyl
solubility enhancement provided by the ternary complex methylcellulose; PVP, polyvinylpyrrolidone. Adapted from Ribeiro et al.
(Mura et al., 2001b). (2003a).
396 Williams et al.

(Jarho et al., 1998). Water-soluble polymers (at low con- 4.04 mM (He et al., 2003), and although similar
centrations), therefore, offer a potentially synergistic increases in solubility are possible via the addition of
benefit for CD-mediated drug solubilization. cosolvent, this is evident only on the addition of large-
The mechanism by which water-soluble polymers volume fractions (.50% cosolvent). In contrast, appro-
enhance drug solubilization in the presence of CDs is priate combinations of 20% HP-b-CD with methanol,
not fully understood (polymers typically have little ethanol, and propanol are all capable of increasing
intrinsic effect on drug solubility), although it has been solubility above 4.04 mM (He et al., 2003). Interest-
suggested that the polymers interact directly with the ingly, at lower CD concentrations (i.e., #10% HP-
drug-CD complex to form a ternary drug-CD-polymer b-CD), low concentrations of cosolvent (typically ,25%
structures with enhanced solubilization capacity (Lofts- cosolvent) decreased or had no impact on drug solu-
son et al., 1994; Loftsson and Brewster, 1996; Messner biliztion (He et al., 2003). The potential for cosolvent to
et al., 2010). In support of this contention, solid-state lower drug solubilization in the presence of lower
analysis of vinpocetine-CD-polymer ternary complexes concentrations of CD likely reflects the competitive
reveals evidence of hydrogen bond interactions be- effects of the cosolvent for the hydrophobic cavity. He
tween the polymer and the drug-CD complex (Ribeiro and coworkers (2003) noted that a higher-molecular-
et al., 2003b). However, other interactions (such as van weight cosolvent (propanol) showed the highest capac-
der Waals or ion-dipole interactions) cannot be dis- ity to lower drug solubility in the presence of CD and
counted (Ribeiro et al., 2005a). Indeed, Valero et al. subsequently proposed that for the lower molecular
(2003) suggest that PVP is able to form ion-dipole in- weight cosolvents, CDs may simultaneously complex
teractions with naproxen sodium while the drug is cosolvent and drug, thereby preventing the cavity com-
enclosed within the hydrophobic cavity of HP-b-CD, and petition effects evident with propanol (see also Section
simultaneously to form hydrogen bonds with the CD. VI.D.3).
Interestingly, in the latter example, the addition of PVP 3. Cyclodextrins and Surfactants. Cyclodextrins
did not enhance CD solubilization of naproxen, and form inclusion complexes with both nonionic (Veiga
Smith et al. (2005) report that PVP is unable to enhance and Ahsan, 1998; Buschmann et al., 2000; Eli et al.,
the solubilization of carbamazepine by SBE-b-CD. 2000) and ionic surfactants (Park and Song, 1989;
Enhancement of CD solubilization via the addition of Junquera et al., 1993), leading to potential competition
water-soluble polymers is therefore not universal. between drug and surfactant for CD binding sites.
Polymers may also stabilize drug-CD aggregates and Furthermore, the complexed surfactant fraction is
reduce the potential for deaggregation on dilution or unable to participate in micelle formation, reducing
agitation (Loftsson et al., 2002; Messner et al., 2011b). micellar solubilization (Junquera et al., 1993). Combi-
Thus, 0.5% carboxymethyl cellulose (a negatively nations of CDs and surfactants therefore often lead to
charged polymer) stabilizes hydrocortisone–HP-b-CD a decrease in the concentration of solubilized drug
aggregates (Messner et al., 2011a), although in this (Veiga and Ahsan, 1998; Tommasini et al., 2004; Yang
case, the addition of 5% of a neutral (HPMC) or posi- et al., 2004; Rao et al., 2006) and are commonly avoided
tively charged (hexadimethrine bromide) additive had (see also Section VII.D.4).
little effect on aggregate stability (Messner et al.,
2011a). E. In Vivo Utility of Drug-Cyclodextrin Complexes
2. Cyclodextrins and Cosolvents. Cosolvents such as in Parenteral Formulations
short-chain alcohols (e.g., ethanol), low-molecular- Cyclodextrins have been widely explored for their
weight polyethylene glycols, and propylene glycol are potential to improve the aqueous solubility of poorly
often used to enhance drug solubility (see Section VI) water-soluble drugs in parenteral formulations. CDs
and may also be used in combination with CDs. approved for use in marketed parenteral products in-
However, cosolvent molecules compete with drug clude SBE-b-CD (Captisol; Ligand Technology, La
molecules for the CD hydrophobic cavity and also Jolla, CA) and HP-b-CD (Table 21). Sporanox IV
promote the solubility of the noncomplexed drug in free (Janssen Pharmaceuticals) contains 1% itraconazole,
solution (Loftsson et al., 1993). As such, drug-CD 40% HP-b-CD, and 2.5% propylene glycol with pH
binding may be reduced (van Stam et al., 1996; Miyake adjustment to pH 4.5. It is diluted 1:1 with saline
et al., 1999b; Evans et al., 2000; Partyka et al., 2001). before administration via i.v. infusion (Strickley, 2004).
Nonetheless, combinations of CDs with cosolvents may Vfend IV (Pfizer) is a freeze-dried mixture of voricona-
lead to net increases in drug solubilization compared zole (0.667%) and SBE-b-CD (10.67%). Geodon (Pfizer)
with cosolvent or CD alone (Reer and Muller, 1993; is also a freeze-dried product containing ziprasidone
Yalkowsky, 1999; Li et al., 1999d; Nandi et al., 2003). mesylate trihydrate (2%) and SBE-b-CD (29.4%). After
Consistent with this suggestion, the marketed itraco- reconstitution, Geodon is administered via i.m. injection.
nazole formulations (Sporanox IV and Oral Solution) 1. Effect of Cyclodextrins on Drug Pharmacokinetics
contain both CD and propylene glycol. Similarly, the after Intravenous Administration. After i.v. adminis-
solubility of fluasterone in a 20% HP-b-CD solution is tration, CD-based formulations are rapidly diluted in
Strategies for Low Drug Solubility 397

plasma and interact with endogenous components in


the blood that competitively displace drug molecules
from the CD binding cavity. Dilution and displacement
therefore promote drug dissociation from CDs, allow-
ing distribution of the drug to sites of action (Rajewski
and Stella, 1996; Stella and Rajewski, 1997; Stella
et al., 1999; Stella and He, 2008).
For drugs with CD-binding constants of up to 2  104
21
M , a 1 in 100 dilution factor has been suggested to
facilitate drug dissociation from CDs (Thompson,
1997). Simply assuming a blood volume of 3.5 L, the
administration of 5 ml of a CD formulation results in
a 1:700 dilution. In practice, the volumes of distribu-
tion of most CDs are closer to the volume of extra-
cellular water (;30% body weight) rather than blood
volume; therefore, even greater dilution factors of up to
1:4200 are likely (Stella et al., 1999). Cholesterol
(Uekama, 2004) and albumin (Frijlink et al., 1991; Fig. 37. Plasma concentrations of OZ209 (expressed as the free base) in
Lin et al., 2007a) compete for binding to the hydropho- Sprague–Dawley rats after a 30-min i.v. infusion of 25 mg/kg (mesylate
salt) in a 0.1 M sulfobutylether-b-cyclodextrin (SBE-b-CD) formulation
bic cavity of CDs and, together with the dilution effects, and a cyclodextrin-free formulation. Values are expressed as means (n 5
promote drug dissociation from the CD cavity under 4) 6 SD. Time zero represents the start of the infusion. Adapted from
Charman et al. (2006).
most circumstances. Consistent with these sugges-
tions, many reports describe no difference in drug
pharmacokinetics following administration as either and CD and also results in a linear decrease in the
a CD complex or in the uncomplexed form (Stella and solubilizing capacity of the system. In most cases,
He, 2008). Examples include isotretinoin (Lin et al., dilution of a 1:1 drug-CD complex is therefore unlikely
2007a), melphalan (Koltun et al., 2010), prednisolone to result in precipitation, regardless of the extent of
(Arimori and Uekama, 1987), propofol (Egan et al., dilution (Rajewski and Stella, 1996; Yalkowsky, 1999).
2003), miconazole (Piel et al., 1999), and etomidate This provides a significant advantage over other solu-
(McIntosh et al., 2004). bilization modalities such as the addition of cosolvent,
In contrast, where drug-CD binding affinity is very where dilution leads to exponential decreases in solu-
high (.1  105 M21), CD complexation may perturb bilization capacity and significantly elevated risks of
drug pharmacokinetics (Charman et al., 2006; McIn- precipitation. In contrast, dilution of a 1:2 drug-CD
tosh et al., 2010). For example, the volume of distri- complex leads to a nonlinear decrease in solubilization
bution and clearance of OZ209 (a novel antimalarial capacity with dilution, introducing an increased risk of
compound) is significantly decreased in rats when drug precipitation after dilution to low CD concen-
dosed as a complex of OZ209 and SBE-b-CD compared trations (Rajewski and Stella, 1996) (Fig. 38B).
with an uncomplexed (pH-buffered solution) formula- 3. Potential Toxicological Properties of Parenterally
tion (Charman et al., 2006) (Fig. 37). The binding Administered Cyclodextrins. The in vivo toxicology of
constant of OZ209 with SBE-b-CD is notably high and parenterally administered CDs has been described in
between 1  105 2 1  106 M21 (Perry et al., 2006). detail elsewhere (Rajewski and Stella, 1996; Irie and
Stella and He (2008) have subsequently proposed Uekama, 1997; Stella and He, 2008) and is therefore
a threshold binding constant of 1  105 M21, above addressed only briefly here.
which dilution on administration may be insufficient The major potential toxicities associated with CD
for complete dissociation of the complex. administration stem from membrane disruption, cho-
A more detailed description of the effect of CDs on lesterol complexation, and precipitation of insoluble
drug pharmacokinetics and distribution after paren- cholesterol-CD complexes in the kidney and commonly
teral administration has been provided by Buggins et al. lead to nephrotoxicity (Irie and Uekama, 1997;
(2007). Uekama, 2004; Stella and He, 2008). The hemolytic
2. Effects of Dilution on Drug Solubilization by properties of a number of chemically modified CDs
Cyclodextrins. Although dilution and displacement toward rabbit erythrocytes have been correlated with
on administration provide a ready mechanism for drug the ability to solubilize cholesterol. In these studies,
liberation from CDs, dilution also increases the risk HP-b-CD and SBE-b-CD were less hemolytic than the
of drug precipitation under some circumstances. The more lipophilic methylated CDs, and SBE-b-CD was
effect of dilution of a solution containing a 1:1 drug-CD less hemolytic than HP-b-CD (Irie and Uekama, 1997;
complex is shown graphically in Fig. 38A. Dilution Uekama, 2004). A 5% w/v SBE-b-CD formulation of
leads to a linear reduction in the concentration of drug etomidate also resulted in less (10-fold) hemolysis on
398 Williams et al.

urine. The elimination half-lives of HP-b-CD and SBE-


b-CD are less than 2 h in humans (;114 min and ;84
min for HP-b-CD and SBE-b-CD, respectively), with
complete elimination within 6 to 12 h (Zhou et al.,
1998; Stella and He, 2008). The rate of clearance of
administered CDs may be lower under conditions of re-
duced glomerular filtration in patients suffering from
renal insufficiency (Irie and Uekama, 1997), potentially
leading to accumulation after repeated administration
(Stella and He, 2008). More detailed information on
maximum possible doses for commonly used CDs is
available in reviews by Brewster and Loftsson (2007)
and Stella and He (2008).

F. In Vivo Utility of Drug-Cyclodextrin Complexes


in Oral (Nonparenteral) Formulations
Cyclodextrins have been widely explored as possible
excipients in oral solution formulations, although
currently, Sporanox Oral Solution is the only marketed
product; the oral bioavailability of itraconazole from
this formulation is estimated to be between 45 and 82%
(Strickley, 2004). CDs may be used as components in
solid oral dosage forms as rapid solvent removal from
drug-CD solutions does not destroy the complex. The
potential for CDs to improve oral bioavailability has
been reviewed elsewhere, and the interested reader is
directed to this reference for greater detail (Carrier
et al., 2007).
1. Effect of Cyclodextrins on Oral Drug Bioavailability.
The ability of CD complexation to enhance the effec-
tive aqueous solubility of poorly water-soluble drugs
Fig. 38. Effect of dilution on drug solubility in cyclodextrin (CD) results in increases in oral bioavailability in many
formulations in instances where drug forms (A) a 1:1 complex with CD
or (B) 1:1 and 1:2 complexes with CD. S0 is the intrinsic drug solubility in cases. For example, formulations containing either
the absence of CD. SBE-b-CD or HP-b-CD increase the oral bioavailabil-
ity of cinnarizine in fasted beagle dogs from 8.0 6 4%
after administration of a simple aqueous suspension
i.v. administration to dogs compared with a commer- formulation to 55 to 60% after administration of
cially available propylene glycol–based formulation the CD formulation (Fig. 39) (Jarvinen et al., 1995).
(Amidate, Abbott Laboratories) (McIntosh et al., The oral bioavailability of resveratrol trimethyl ether
2004). The potential for hemolysis after administration (trans-3,5,49-trimethoxystilbene) is also increased
of HP-b-CD and SBE-b-CD, therefore, appears to be from ,1% after administration of an oral suspension
no greater than, and likely less than, many other formulation (to rats), to 64.6 6 8.0% after administra-
solubilization strategies. tion of a solution containing 0.3 M randomly methylated
The renal toxicity of CDs is highly linked to CD b-CD (RM-b-CD) (Lin and Ho, 2011). In the same
solubility. Thus, i.v. administration of the relatively study, dose escalation from 15 mg/kg to 60 mg/kg led
poorly water-soluble b-CD to rats at ;0.9 g/kg dose to corresponding increases in exposure (Lin and Ho,
leads to significant renal toxicity and some irritation at 2011). Melarsoprol is a poorly water-soluble drug (aque-
the site of administration (Frank et al., 1976; Irie and ous solubility 6 mg/ml) used in the treatment of human
Uekama, 1997; Thompson, 1997). The renal toxicity of African trypanosomiasis (HAT) (Gibaud et al., 2005).
natural CDs follows the rank order b-CD (most toxic) . Because of its low water solubility, melarsoprol is
a-CD . g-CD (Thompson, 1997), a trend consistent administered i.v. as a cosolvent formulation in the
with solubility where g-CD (most soluble) . a-CD . commercially available product (Arsobal; Sanofi-Aventis).
b-CD (Irie and Uekama, 1997). The low aqueous solu- Complexation with RM-b-CD or HP-b-CD, however,
bility and subsequent nephrotoxicity of b-CD preclude increases melarsoprol solubility by more than 7000-
use in parenteral formulations. In contrast, modified fold (Gibaud et al., 2005), and oral administration of
CDs are more soluble in water, show lower renal a reconstituted lyophilised drug-CD complex (RM-
toxicities, and are usually eliminated unchanged in the b-CD or HP-b-CD) over a 7-day period eliminated the
Strategies for Low Drug Solubility 399

Alternatively, spray-drying or free-drying drug-CD


solutions yield powders that can contain solid drug-CD
complexes, and formulations containing such pre-
formed CD complexes typically result in more signif-
icant gains in dissolution and oral bioavailability
compared with physical mixtures (Fig. 40) (Corrigan
and Stanley, 1982; Savolainen et al., 1998; Okimoto
et al., 1999; Nagarsenker et al., 2000; Fernandes et al.,
2002; Koester et al., 2003; Carrier et al., 2007; Fukuda
et al., 2008). Techniques and approaches used to verify
that drug-CD complexes are retained in the solid state
are discussed in more detail in Section VIII.G.2.
3. Using Cyclodextrins to Stabilize Amorphous Drug.
In the solid state, drug molecules complexed within the
hydrophobic cavity of a CD molecule are constrained
from interaction with other drug molecules and are
Fig. 39. Cinnarizine plasma concentration in dogs after oral adminis- typically unable to interact sufficiently to form a crystal
tration of 25 mg cinnarizine in either sulfobutylether-b-cyclodextrin
(SBE-b-CD) or hydroxypropyl-b-cyclodextrin (HP-b-CD) solutions, a cin- lattice (Bettinetti et al., 1992; Mura et al., 1998; Szejtli,
narizine aqueous suspension (pH 4.5), or capsules containing cinnarizine 1998; Kimura et al., 1999). This mechanism for
with and without SBE-b-CD. Values are expressed as means (n 5 4) 6
SEM. Adapted from Jarvinen et al. (1995). stabilizing drug in the noncrystalline state is in some
ways analogous to mesoporous systems, where crystal
growth is limited by the narrow dimensions of the
parasite from HAT-infected animals (Rodgers et al., inner capillary network (see Section XII.A). Modified
2011). Capsules containing lyophilized ziprasidone CDs are also commonly present in the amorphous
mesylate and SBE-b-CD complexes and HPMC acetate form, and solid drug-CD inclusion complexes are
succinate (HPMCAS) dosed to fasted humans also therefore often isolated in the amorphous or micro-
showed increased oral bioavailability compared with crystalline form (Uekama et al., 1983; Veiga et al.,
commercial Geodon capsules (containing ziprasidone 1996; Hirayama et al., 1997; Mura et al., 1998; Veiga
hydrochloride, lactose, pregelatinized starch and mag- et al., 2001; Sotthivirat et al., 2007). As such, CD
nesium stearate) and a diminished food effect (Thom- complexation is an alternate approach to isolate drug
bre et al., 2011). Paclitaxel-CD complexes (HP-b-CD or in a noncrystalline amorphous form, and solid-state
b-CD) encapsulated within poly(anhydride) nanopar- CD complexes may be regarded in a broad sense as
ticles also facilitate increased oral bioavailability for a class of solid dispersion. The potential benefits of
paclitaxel compared with the commercial Taxol formu- isolating a drug in the amorphous form to drug
lation. The paclitaxel-CD nanoparticle formulations dissolution and bioavailability are described in more
resulted in absolute bioavailability of 80% relative to detail in Section X. In brief, however, the loss of
i.v. administration of Taxol. In contrast, oral adminis- crystalline character reduces the barrier to solubility
tration of the Taxol formulation did not provide plasma presented by solute-solute molecular interactions in
levels above the limit of quantification of the assay the crystal matrix and this change to solid-state
(Agueros et al., 2010). The authors suggested that the properties of the drug leads to enhanced dissolution
dramatic increase in oral bioavailability reflected in- rates. In the case of a CD-based matrix, dissolution of
creases in paclitaxel solubility mediated by the CDs and the carrier also increases the solubilization capacity of
bioadhesion resulting from the nanoparticle formulation. the dissolution media, further enhancing the drivers of
2. Oral Administration of Physical Mixtures Versus dissolution rate (Fig. 40).
Isolated Drug-Cyclodextrin Complexes. The simplest As with all formulations containing amorphous
drug-CD formulations are obtained by blending drug (i.e., metastable) drug, rapid dissolution of the amor-
and CD to form a physical powder mixture that can be phous form may result in drug concentrations in the
blended with other excipients and compressed to form dissolution media that are supersaturated with respect
traditional solid dosage forms (see Section VIII.G.2). to the solubility of the stable crystal. Supersaturation
Drug-CD physical mixtures improve the rate of drug may be beneficial with respect to drug absorption but is
dissolution over that of drug alone, suggesting that also a driver of drug precipitation. The net benefit is
CDs can enhance drug dissolution via either improved therefore a balance between increased dissolution rate,
wetting (Guyot et al., 1995; Nagarsenker et al., 2000; enhanced supersaturation, and protection against pre-
Reddy et al., 2004; Shinde et al., 2008) or in situ cipitation. CDs, including HP-b-CD and SBE-b-CD,
formation of the drug-CD complex in the dissolution enhance the stability of supersaturated drug solutions
medium (Corrigan and Stanley, 1982; Guyot et al., (Vandecruys et al., 2007; Brewster et al., 2008a,b), sug-
1995; Fukuda et al., 2008). gesting that CDs are able to both promote the isolation
400 Williams et al.

A number of studies have examined the effect of


adding polymers to CD formulations. In most cases,
however, the intent has been to exploit the capacity
of polymers to enhance drug-CD complexation (see
Section VIII.D.1) (Loftsson et al., 1994; Smith et al.,
2005) or to generate controlled release CD-based dosage
forms (Uekama et al., 1990; Rao et al., 2001; Koester
et al., 2003, 2004; Ribeiro et al., 2005b) rather than to
stabilize the supersaturated solutions that may result
on dissolution (Brewster et al., 2002; Lee et al., 2009).
Consistent with other formulation approaches that
contain amorphous solids, CD-based amorphous dis-
persions are thermodynamically unstable. Amorphous
drug contained within CD solid complexes is therefore
expected, over time, to revert to the more stable but
less soluble crystalline form. Unfortunately, whether
the drug-CD complex slows the process of recrystalli-
zation and whether this is dependent on the affinity of
the drug-CD interaction is not well described, and
compared with polymer-based solid dispersions, rela-
tively few studies have investigated the stability of
amorphous drug-CD formulations (Irie and Uekama,
1997; Hirayama et al., 2001; Uekama, 2002). However,
given that the intermolecular interactions between drug
and CD must be overcome for the drug to crystallize, it is
conceivable that a correlation exists between the
physical stability of the amorphous drug-CD complex
and the strength of the interaction between drug and
CD. This is comparable with the physical stability
within amorphous solid dispersions, where evidence of
intermolecular hydrogen bonds between drug and poly-
mer is often considered beneficial to the long-term
physical stability of the formulation.
4. Cyclodextrin Complexation as a Limitation to Oral
Bioavailability. Enhanced drug absorption after oral
administration of CD formulations is commonly as-
Fig. 40. (A) Release profiles of ketoprofen in 900 ml of 0.1 N HCl and 0.1 cribed to increases in the apparent solubility of the
N HCl containing either 25mg of sulfobutylether-b-cyclodextrin (SBE- drug in the GI fluids (resulting from either complex
b-CD) or 25 mg of b-cyclodextrin (b-CD). (B) Effect of hot-melt extrusion
processing with either SBE-b-CD or b-CD on ketoprofen release in 900 ml formation, isolation of the amorphous form, or super-
of 0.1 N HCl. Experiments carried out at 37 6 0.5°C USP apparatus 2, 50 saturation stabilization). However, there are also
rpm. Values are expressed as means (n 5 6) 6 SD. Adapted from Fukuda
et al. (2008).
several examples where CDs have increased drug
solubility but have not improved oral bioavailability
(Nambu et al., 1978; Oteroespinar et al., 1991; Miyake
of the amorphous form and stabilize the supersatu- et al., 2000). Additional factors must therefore be
rated solutions that form on dissolution. Supersatura- considered when attempting to understand the perfor-
tion stabilization likely reflects increases in apparent mance of CD formulations in vivo (Jambhekar et al.,
drug solubility in the presence of CD and a reduction in 2004; Uekama, 2004; Carrier et al., 2007). In particu-
the degree of supersaturation relative to the solubility lar, drug association with a CD complex reduces
of the stable crystal form. Alternatively, and in a thermodynamic activity compared with a simple solu-
manner analogous to the use of polymers in solid tion formulation, and for drugs with high CD binding
dispersions or lipid-based formulations (Brewster constants, this may be sufficient to reduce the rate or
et al., 2002; Chowdary and Srinivas, 2006; Cappello extent of drug absorption. In a similar fashion to drug
et al., 2007; Lee et al., 2009; Warren et al., 2010; Anby dissociation from the complex after i.v. administration,
et al., 2012), CDs may stabilize supersaturation by drug dissociation in the GI tract is expected to be
slowing/inhibiting nucleation or crystal growth (Brewster driven by dilution and competition for CD complexa-
et al., 2008a). tion by endogenous substances in the GI lumen, such
Strategies for Low Drug Solubility 401

as bile acids and cholesterol (Nakanishi et al., 1989; natural and chemically modified CDs has been exten-
Rajewski and Stella, 1996; Miyake et al., 1999b). sively reviewed elsewhere (Irie and Uekama, 1997;
In the case of the cinnarizine and HP-b-CD or Thompson, 1997; Uekama et al., 1998; Stella and He,
SBE-b-CD complexes described previously, an increase 2008), and in general, CDs are less toxic after oral
in Cmax and oral bioavailability and decrease in Tmax administration compared with i.v. administration since
compared with a suspension formulation suggest effi- oral bioavailability of CDs is low; natural a-CDs and
cient and rapid dissociation of cinnarizine from the b-CDs are not substrates for salivary and pancreatic
HP-b-CD or SBE-b-CD complex in vivo (Jarvinen et al., amylases and are also resistant to acid hydrolysis in
1995). For cinnarizine, however, drug-CD binding the stomach. Conversely, g-CD is rapidly digested into
constants are not high (,5000 M21) (Jarvinen et al., glucose and maltose by amylases in the GI tract (De
1995), whereas poor dissociation and the potential for Bie et al., 1998). All CD varieties may be subject to
limitations to drug availability may be more likely for fermentation by microflora present in the large in-
drugs with higher binding affinity (Tokumura et al., testine (Flourie et al., 1993), although chemical sub-
1986; Bekers et al., 1991; Uekama et al., 1998; Stella stitution is thought to reduce the extent of breakdown
et al., 1999; Carrier et al., 2007). Indeed, several (Irie and Uekama, 1997).
studies have shown a decrease in in vitro membrane The high molecular weight (usually .1000 Da) and
flux for drugs solubilized by CDs compared with sol- hydrophilicity of CDs limits membrane permeability;
ution formulations alone (Iervolino et al., 2000; Dias therefore, absorption after oral administration is low
et al., 2003; Manca et al., 2005; Dahan et al., 2010), (Irie and Uekama, 1997; Stella and He, 2008). Indeed,
consistent with a decrease in thermodynamic activity less than 1% of an orally administered dose of HP-
and similar, for example, to the reduction in free con- b-CD is absorbed in rats (Stevens, 1999), and similarly
centration that results from micellar solubilization. low (0.5–3.3%) absorption of HP-b-CD has been re-
A relatively complex series of potentially conflicting ported in humans (Thompson, 1997; Brewster and
processes is therefore evident during the dissolution of Loftsson, 2007). Much less oral safety data are avail-
a CD-based formulation compared with a suspension able for SBE-b–CD; however, absorption is likely to be
formulation. The CD formulation is expected to lead to low (consistent with HP-b-CD), with most ingested CD
a significant improvement in the dissolution rate of the excreted via the feces (Thompson, 1997). Predictably,
drug-CD complex, but subsequent dissociation of drug more lipophilic CD derivatives, including randomly
from the complex is required before absorption [as only methylated (RM) and dimethylated (DM) b–CD deriv-
noncomplexed (free) drug is absorbed]. The extent of atives, are absorbed to a greater extent, although oral
dissociation reflects the magnitude of the drug-CD bioavailabilities remain relatively low. For example, in
binding constant, and, ultimately, whether the rats, ;10% of DM-b–CD (Szatmari and Vargay, 1988;
strength of the association complex limits bioavailabil- Thompson, 1997) and ;12% of RM-b–CD (Antlsperger
ity will reflect a balance between the degree of and Schmid, 1996) are bioavailable after oral admin-
solubility enhancement and the magnitude of the istration.
reduction in thermodynamic activity. The impact of Cyclodextrins have been shown to complex with bile
complexation on bioavailability will also depend on salts that are present in the intestinal lumen (Tan and
drug permeability across the intestinal membrane, Lindenbaum, 1991; Holm et al., 2011). Rats repeatedly
where increased permeability is expected to provide dosed orally with 5 g/kg HP-b-CD developed pancreatic
improved sink conditions and promote continuing hypertrophy (and later, pancreatic neoplasia), and this
dissociation of the drug-CD complex. Accordingly, was attributed to bile salt-CD complexation in the
CDs have, in general, a greater capacity to improve intestine and increased fecal excretion of bile salts, in
the oral absorption of drugs belonging to class II of the turn promoting the release of cholecystokinin, a peptide
BCS (i.e., low-solubility, high-permeability drugs) with known mitogenic properties in rats (Gould and
(Loftsson et al., 2007b; Loftsson and Brewster, 2010). Scott, 2005). However, as cholecystokinin shows no
Nonetheless, improved oral absorption of BCS class IV mitogenic effects in humans and other species, the
drugs (i.e., those of low solubility and low permeability) effects of HP-b-CD on the pancreas may be specific to
is also possible, suggesting that bioavailability enhance- the rat (Gould and Scott, 2005).
ment via increases in solubility are possible even for low Similar to parenterally administered CDs, those that
permeability compounds (Loftsson and Brewster, 2010). enter the systemic circulation via the GI tract are
Formulation design should ensure that the CD concen- largely excreted through the kidneys. The combination
trations used are not higher than that required to provide of low absorption and the relatively low systemic
adequate solubilization since surplus CD will decrease the toxicity of HP-b-CD and SBE-b-CD dictates that large
free drug concentration thereby reducing flux (Dahan quantities may be tolerated after oral administration,
et al., 2010; Loftsson and Brewster, 2011). with doses of up to 8 g per day of HP-b-CD used in the
5. Potential Toxicological Properties of Orally Ad- clinic (Gould and Scott, 2005; Loftsson and Brewster,
ministered Cyclodextrins. The oral toxicology of both 2010). No marketed products for oral administration at
402 Williams et al.

present contain SBE-b-CD, although doses via i.v. vacuum to remove the solvent (Arima et al., 2001b;
administration may be as high as 14 g per day Mura et al., 2001a; Patel and Vavia, 2006; Veiga et al.,
(i.e., Vfend IV). 1996). More commonly, however, solid complexes are
6. Effect of Cyclodextrins on Membrane Transport isolated from freeze-dried or spray-dried solutions
and Drug Metabolism. Several studies have sug- (Marques et al., 1990; Jarvinen et al., 1995; Veiga
gested that CDs, in particular the modified CDs, may et al., 1996; Savolainen et al., 1998; Dollo et al., 1999;
affect oral drug absorption via effects on intestinal Jambhekar et al., 2004; Pralhad and Rajendrakumar,
membrane transporter functionality (e.g., via effects on 2004; Villaverde et al., 2004) or after coprecipitation
the efflux transporter, P-gp) or via alterations in (Uekama et al., 1982; Badawy et al., 1996; Veiga et al.,
cytochrome P450–based first-pass metabolism (Arima 1996). In the latter approach, solid drug-CD complexes
et al., 2001a, 2004; Yunomae et al., 2003; Ishikawa are encouraged to precipitate from a supersaturated
et al., 2005; Fenyvesi et al., 2008). These effects are drug-CD solution. Where the aqueous solubility of the
analogous to the potential effects of surfactants on inclusion complex is low, water may be used as the
drug efflux and intestinal metabolism (Friche et al., solvent; however, small amounts of cosolvent are often
1990b; Kemper et al., 2003; Buggins et al., 2007; Yang added where the drug-CD solubility is high (as is often
et al., 2007), however, the mechanisms by which CDs the case). The solvent is then removed by evaporation
inhibit efflux or metabolism are less well described but to stimulate precipitation of the complex (Kou et al.,
presumably reflect complexation of membrane lipids, 2011).
reorganization of membrane structures, or alterations Similar to the manufacture of nano and micro-
to transporter or enzyme activities (Arima et al., particulates (see Section IX) and solid dispersions
2001a, 2004; Ishikawa et al., 2005). Also consistent (see Section X), supercritical fluid (SCF) techniques
with surfactants, definitive data describing the effect of may be used in the manufacture of solid drug-CDs
CDs on transporter or enzyme functionality are formulations. In these techniques, SCFs may be used
difficult to obtain in vivo for poorly water-soluble drugs as either solvent or antisolvent (Van Hees et al., 1999;
since bioavailability is likely to be enhanced by both Junco et al., 2002; Al-Marzouqi et al., 2006); however,
changes to dissolution and efflux/metabolism. Delinea- use as an antisolvent is more common as the solubility
tion of the two pathways is therefore difficult without of many drugs and CDs in SCFs is low, limiting the
recourse to knockout animals, the latter restricting drug-CD concentrations that may be used (Tozuka
studies to small rodent models. et al., 2006).
Solid complexes should be analyzed to verify the
G. Manufacture of Cyclodextrin Formulations physical form of the drug and to ascertain whether the
1. Parenteral Formulations. As discussed in Section integrity of the drug-CD inclusion complex has been
VIII.D, CDs are widely used in conjunction with other maintained during processing (Pedersen, 1997; Thom-
formulation strategies during the development of bre et al., 2012). Loss of drug crystalline character after
parenteral formulations as the combined use of CDs precipitation/drying of drug-CD solutions and the
with, for example, pH adjustment may lead to maintenance of drug crystallinity in physical mixtures
synergistic increases in drug solubilization. Conse- of drug and CD are often taken as evidence of the
quently, CD formulations typically contain several stability of drug-CD complexes during the manufacture
different components, including polymers, buffers, of solid material. However, many of the common
cosolvents, and such. Parenteral formulations are characterization techniques used to make these dis-
usually prepared as simple aqueous solutions but tinctions, including DSC and XRPD, lack the capacity
may also be freeze-dried to form solid inclusion com- to discriminate between complexed and uncomplexed
plexes (Brewster et al., 1991; Xiang and Anderson, amorphous drug. Spectroscopic methods such as
2002; Lockwood et al., 2003; Hong et al., 2011) that are Raman or FTIR provide a more robust means for
reconstituted with water or saline immediately before characterizing the nature of drug-CD complexes in the
administration (Kim et al., 2004a; Strickley, 2004). The solid state (Veiga et al., 1996; Van Hees et al., 1999;
quantity of CD required to dissolve drug to achieve Fernandes et al., 2002; Jambhekar et al., 2004) since
a particular target concentration is determined by spectral shifts indicate the formation of drug-CD in-
phase solubility analysis or, where published binding teractions, while the absence of spectral bands specific
constant information is available, may be calculated to the drug provide evidence that the drug is shielded
using eq. 41 (see Section VIII.B.1). from the spectroscopic signal and is therefore most
2. Solid Drug-Cyclodextrin Complexes. Solid drug- likely located within the CD cavity (Jambhekar et al.,
CD complexes are generated most simply by tritura- 2004).
tion, during which a small amount of solvent (usually
a mixture of organic solvent with water) is added to H. Summary
a mixture of drug and CD to form a slurry of paste-like Cyclodextrins are cyclic oligosaccharides that are
consistency. The slurry is subsequently dried under a able to form guest-host inclusion complexes with poorly
Strategies for Low Drug Solubility 403

water-soluble drugs. The affinity of drug molecules for of nanoscale drug assemblies from highly supersatu-
solubilization in the hydrophobic core of the CD is rated drug solutions.
defined by the CD binding constant, and, for most The term nanoparticle is frequently used to describe
drugs, binding constants of 1  102 2 1  104 M21 are a range of nanosized materials, most commonly poly-
evident. Increases in the drug solubility of several meric, liposomal, or solid lipid nanoparticles, but it also
orders of magnitude are possible, and solubilization correctly describes nanoparticulate drug crystals and
efficiency may be enhanced by coformulation with suspensions (Wu et al., 2011). Drug nanoparticles have
polymers or cosolvents and via combination with pH many advantages over simple micronized drug pow-
manipulation strategies. Coformulation with surfac- ders. Indeed, the increase in surface area to mass ratio
tants is typically detrimental to solubilization. CDs for nanoparticulates is dramatic, sometimes covering
have been used to facilitate the generation of oral and several orders of magnitude (Fig. 41) (Junghanns and
parenteral formulations and as simple solutions or as Muller, 2008; Merisko-Liversidge and Liversidge, 2008),
solid or reconstitutable solid dosage forms. Solid CD- and this has the potential to provide for substantial
drug complexes may also be used to isolate drug in increases in dissolution rate, bearing in mind the de-
the more soluble amorphous form, although detailed pendence of dissolution rate on surface area (eq. 9).
studies of the long-term physical stability of amor- Particles in the nanoscale may also be more soluble
phous drug-CD complexes are limited. The toxicity of because of changes to particle curvature and the intro-
modified CDs appears to be similar or lower than that duction of defects into the crystal lattice (Junghanns
of common surfactants or cosolvents and is reduced and Muller, 2008; Mauludin et al., 2009). As a result,
after oral administration as a result of the limited
drug nanocrystals may yield oral bioavailabilities that
absorption of the excipient.
far exceed those obtained using conventional particles
or micronized particles of poorly water-soluble drugs
IX. Particle Size Reduction Strategies (Liversidge and Cundy, 1995; Jinno et al., 2006; Sun
et al., 2011). The practical applicability of nanocrystals
Particle size reduction leads to an increase in the
is evident in several marketed nanocrystal formulations
surface area available for solvation and an increase in
(Table 23) (Chaubal, 2004; Rabinow, 2004; Merisko-
the rate of dissolution for solid drug products. Particle
size reduction technologies are therefore routinely used Liversidge and Liversidge, 2011), and there is also
to improve the oral bioavailability of poorly water- increasing application of nanosuspensions in preclinical
soluble drugs (Rabinow, 2004; Keck and Muller, 2006; studies, where the ability to enhance dissolution at the
Merisko-Liversidge and Liversidge, 2011) and when same time as maximizing drug dose (since nanocrystals
developing fine (nano) suspension formulations for pa- require little additional excipient) is often critical.
renteral delivery (Muller and Bohm, 1998; Shi et al., Importantly, the development of top-down nanomilling
2009). technologies now allows the processing of drug batches
Particle size reduction technologies may be charac- relatively cheaply and in batch sizes suitable for appli-
terized as either “top-down” processes, where larger cation from early discovery phases (milligram) through
particles are fragmented into smaller particles, or as to product marketing (kilogram) (Chen et al., 2006;
“bottom-up” processes, where small particles are har- Kesisoglou et al., 2007; Niwa et al., 2011b).
vested after recrystallization of a drug from a super-
saturated solution. Traditionally, small particles have
been obtained via top-down, dry-impact processes that
introduce considerable shear forces and reduce the
particle size of a coarse drug powder using, for ex-
ample, hammer mills, ball mills, and air-jet mills. This
process, commonly referred to as micronization, results
in the formation of particles in the micron size range
with average diameters ,10 mm and particle sizes
commonly between 2 and 5 mm (Nykamp et al., 2002;
Rasenack and Muller, 2002). Dry milling remains in
wide use; however, a combination of recent technolog-
ical developments and the wider utilization of poly-
meric and surfactant stabilizers have led to the ability
to produce much smaller (nano) particulates with sizes
in the 200–500-nm range. Bottom-up particle assembly Fig. 41. The change in surface area obtained when solid particles exist in
processes have also become more widespread as cry- the micron-size range (microparticle) and the nanometer-size range
(nanoparticles). Both microparticles and nanoparticles can be prepared
stallization methods have become more sophisticated, from top-down and bottom-up techniques. Adapted from Merisko-Liver-
allowing the controlled crystallization or precipitation sidge and Liversidge (2008).
404 Williams et al.

TABLE 23
List of currently marketed medicines and those under development that contain nanosized drug
Drug Brand Name Technology Route of Administration and Form

Rapamycin Rapamune (Wyeth) NanoCrystal Oral, tablet


Aprepitant Emend (Merck) NanoCrystal Oral, capsule
Fenofibrate Tricor 145 (Abbott) NanoCrystal Oral, tablet
Megestrol MegaceES (Par Pharmaceutical Companies) NanoCrystal Oral, liquid solution
Fenofibrate Triglide (Sciele Pharma Inc) DissoCubes Oral, tablet
Paliperidone palmitate Invega Sustenna (Janssen Pharmaceuticals) NanoCrystal i.v., sustained release injection
Thymectacina Theralux (Celmed) NanoCrystal i.v.
a
Currently in phase I/II clinical trials.

The formation of nanocrystal formulations is not significant (Fig. 41) and may lead to large improve-
without complexity, and several challenges are evi- ments in dissolution rate compared with both un-
dent. Nanoparticles are highly cohesive and must be modified drug powders and micronized drug. Indeed, in
stabilized to prevent aggregation. In addition, not all some cases, dissolution is sufficiently fast as to be
drug candidates have physicochemical properties suit- practically instantaneous [e.g., .80% release within 2
able for effective particle size reduction, and nano- min has been reported for nanocrystals of ibuprofen
particulate drugs may be susceptible to polymorphic (Plakkot et al., 2011) and indomethacin (Liu et al.,
transition on manufacture or storage (Van Eerden- 2011; Niwa et al., 2011a)]. Figure 42 provides a recent
brugh et al., 2008; Wu et al., 2011). The current section illustration of the effect of particle size reduction on
of this review first covers the theoretical background the dissolution of suspensions of itraconazole. Here,
that underpins the effect of particle size on dissolution micronization (to 5.5 mm) leads to a 2- to 3-fold increase
rate, solubility and ultimately in vivo exposure for in dissolution rate, whereas further particle size
poorly water-soluble drugs. Second, the most widely reduction to produce nanoscale particulates (350–700
used methods for nanoparticle production are intro- nm) elicits additional 3- to 5-fold increases in dissolu-
duced. Finally, examples of the potential utility of tion rate compared with the micronized product (and
microparticulate and nanoparticulate formulations af- therefore an overall ;14-fold increase in dissolution
ter oral and parenteral administration are discussed. compared with nonmicronized drug) (Sun et al., 2011).
Although increases in surface area are widely
A. Particle Size Effects on Dissolution, Solubility, and acknowledged as the primary driver of increased dis-
In Vivo Performance solution for microparticulates and nanoparticulates, it
1. Effect of Particle Size on Dissolution Rate. is also apparent that the improvements in dissolution
The rate of drug dissolution from a solid dosage form velocity may exceed that predicted simply on the basis
surrounded by an unstirred diffusional layer is tradi- of available surface area. Under these circumstances,
tionally described by the Noyes-Whitney equation (eq. 9). changes to drug solubility (Cs), to the thickness of the
Where particle size reduction yields an increase in diffusional layer (h), and to particle shape may also
surface area, eq. 9 (see Section I.E) suggests an equi-
valent increase in dissolution velocity and, therefore, an
increase in bioavailability for drugs where exposure
after oral administration is limited by dissolution rate.
Consistent with this suggestion, micronization leads to
an improvement in the oral bioavailability of many
poorly water-soluble drugs, including cilostazol (Jinno
et al., 2006), fenofibrate (Vogt et al., 2008), progesterone
(Hargrove et al., 1989), proquazone (Nimmerfall and
Rosenthaler, 1980), and nitrofurantoin (Watari et al.,
1983), all of which are believed to exhibit dissolution-
rate limited absorption. However, the generation of
micron-sized drug particles is often insufficient to over-
come the challenges to absorption presented by com-
pounds with very low aqueous solubility (,1 mg/ml)
(Kondo et al., 1993; Muller et al., 2001; Keck and Muller,
2006; Gao et al., 2008), and alternative methods have
been developed to reduce particle size further and to
isolate particles with nanometer-scale dimensions.
Fig. 42. In vitro dissolution profiles of itraconazole from the suspensions
The increase in surface area that may be achieved by of various particle sizes in 0.1 M HCl at 37°C. Values are expressed
the generation of nanoparticulate drug suspensions is means (n 5 3). Adapted from Sun et al. (2011)
Strategies for Low Drug Solubility 405

explain the increase in dissolution rate, and these this case of nanoparticulates, however, the degree of
factors are addressed in the sections below. (Anderberg supersaturation is usually low since particle size
et al., 1988; Bisrat and Nystrom, 1988, 1998; Elamin effects on saturated solubility are moderate (,2-fold
et al., 1994; Mosharraf and Nystrom, 1995). owing to changes to curvature or lattice defects). As
2. Effect of Particle Size on Saturated Solubility. a result, the more prominent effects of nanosizing on
Traditional solubility theory suggests that decreases drug absorption are typically driven by the much
in particle size are expected to have little impact on larger changes to dissolution rate.
equilibrium solubility since particle size changes typi- Accurate measurement of the saturated solubility of
cally affect neither solid-state properties nor the effi- a nanoparticulate is also complicated by the potential
ciency of solvation. However, the advent of technologies for very small particles to remain suspended in solu-
that support the isolation of particles with submicron tion after conventional centrifugation (or even ultra-
particle sizes requires that this view be modified. centrifugation in some cases) and to pass through
Indeed, the Ostwald-Freundlich Equation (Freundlich, many of the membrane filters used to separate drug in
1909) (eq. 50), suggests solubility increases on reducing solution from solid drug (Van Eerdenbrugh et al.,
particle size, with effects becoming significant below 1 2010b). Bearing in mind the low intrinsic solubility of
mm, (r 5 0.5 mm): many of the drugs to which this technology is applied,
only trace quantities of particle contamination are req-
Cs 2sVm uired to generate large errors in solubility assessment.
log 5 ð50Þ Recent data to suggest the potential for solid nano-
C‘ 2:303RTrr
crystals to absorb in the UV/Vis light spectrum (Van
where Cs is the saturated solubility, C‘ is the solubility Eerdenbrugh et al., 2011) further reinforce the sugges-
of the solid consisting of large particles, s is the tion that great care must be taken to exclude the
interfacial tension of the solid substance, Vm is the possibility of particulate contamination during solubil-
molar volume of the particles, R is the gas constant, T ity and dissolution assessment.
is the absolute temperature, r is density of the solid, To overcome some of the concerns regarding phase-
and r is the particle radius. separation techniques, light scattering and turbidity
This phenomenon was initially described by Kelvin measurements have been explored as an alternate
in a gas-liquid system, where increasing curvature of approach to solubility assessment (Lindfors et al., 2006;
a liquid droplet (in response to decreasing radius) Van Eerdenbrugh et al., 2010b). In these methods, the
increased vapor pressure and the rate of transfer into first appearance of solid material (as an indication of
a gas phase (described by Wu and Nancollas, 1998). the solubility limit) is determined from the intersection
The principle, however, applies equally to solid parti- of two curves, one describing the small increases in
cles in a liquid medium where a reduction in particle scattering/turbidity (at 700 nm) that are evident
size below 1 mm increases solvation pressure, sub- in undersaturated solutions containing progressively
sequently giving rise to an increase in solubility higher amounts of dissolved drug and a second curve
(Junghanns and Muller, 2008). that describes the relationship between scattering/
In addition to Ostwald-Freundlich effects on curva- turbidity and particle load in a series of saturated
ture, high-energy impact milling and other top-down solutions containing increasing amounts of undissolved
particle size reduction techniques may cause small drug (Van Eerdenbrugh et al., 2010b). In these studies,
defects in the crystal lattice (particularly on the par- correction for scattering in the absence of particulate
ticle surface). This is expected to increase solubility via drug was required as all solutions/suspensions con-
a reduction in the limitations to solubility imparted by tained a small quantity of vitamin E TPGS (as a
the strength of solute-solute interactions in the crystal nanosuspension stabilizer) that resulted in scattering
lattice. from surfactant micelles.
Although changes to kinetic solubility are evident for 3. Effect of Particle Size and Shape on the Diffusional
nanoparticulate drug suspensions via increased curva- Layer Thickness. Particles smaller than 2 mm in
ture (via the Ostwald-Freundlich effect) and the in- diameter demonstrate faster rates of dissolution by
troduction of defects into the crystal lattice, the virtue of a reduction in the thickness of the unstirred
equilibrium solubility of the stable crystal remains diffusional layer (Higuchi and Hiestand, 1963; Lu
unchanged. As such, rapid dissolution of drug nano- et al., 1993). The size of the diffusion layer is described
particles leads to the generation of a transiently super- by the Prandtl equation (eq. 51), where hH signifies the
saturated solution (with respect to the solubility of the thickness of the hydrodynamic boundary layer (the
stable crystal form) and, therefore, the potential for diffusion layer in the Noyes-Whitney equation; eq. 9)
recrystallization. The same situation occurs on disso- around a particle of length (diameter) L over which
lution of all nonequilibrium crystal structures such as a liquid flows at constant velocity V. k is a constant
high-energy polymorphs, solvates, cocrystals (see Sec- (Mosharraf and Nystrom, 1995; Muller and Bohm,
tion V), and, depending on pH, salts (see Section IV). In 1998):
406 Williams et al.
 1=2 
L as this attenuates the increase in free energy associ-
hH 5k × 1=2 ð51Þ ated with the creation of new particle surface and
V
prevents nanocrystal aggregation. Stabilizers are usu-
The Prandtl equation predicts a decrease in hH with ally added before milling initiation, however, recent
decreasing particle size, and from eq. 9, an increase in studies suggest that the dynamic addition of stabilizers
the dissolution rate. The velocity (V) of fluid flow over during milling may further facilitate the milling pro-
a particle in eq. 51 is dependent on particle shape, and cess (Bhakay et al., 2011).
notably slower rates of dissolution can occur where The NanoCrystal technology platform (Elan Drug
smaller particles are irregularly shaped and fluid flow is Technologies, Dublin, Ireland) is the most well de-
slower (Mosharraf and Nystrom, 1995; Sugano, 2008). veloped pearl-milling method for particle size reduc-
tion, although it is available only through a contractual
B. Common Methods to Reduce Particle Size agreement with Elan (recently merged with Alkermes,
Traditionally, micronized drug product has been Dublin, Ireland). This technology utilizes a patented
obtained using dry-impact milling, with the most milling medium (PollyMill) consisting of small (,1
popular equipment including hammer mills, ball mills, mm) highly cross-linked polystyrene beads (or
and more recently, air-jet mills (Hajratwala, 1982; pearls), the design of which facilitates the grinding
Nykamp et al., 2002; Nakach et al., 2004). In these process by maximizing the number of bead-drug
techniques, smaller particles are produced via the particle contact points (Merisko-Liversidge and Liv-
fracture of larger particles and the gradual attrition of ersidge, 2011). The NanoCrystal technology is scale-
the particle surface (Tavares, 2007). However, unless able from small-scale milligram batches (with
very long milling times are used, these techniques are application in early preclinical studies) to large-
unable to produce drug particles with average diame- scale batches (500–1000 kg) that can be progressed
ters in the nanoscale. Particles are also prone to phase to clinical use and marketing. Marketed products that
transformations during dry milling, especially over use the NanoCrystal technology platform cover
extended processing times (Haleblian and McCrone, a range of applications and are listed in Table 23.
1969; Elamin et al., 1994; Mackin et al., 2002; Descamps Although the NanoCrystal technology has domi-
et al., 2007), precluding application to thermally labile nated pearl milling applications for some time, alter-
and low melting materials. native pearl mill technologies have been developed de
In contrast, wet-milling approaches, such as pearl novo (e.g., Dyno-Mill; Willy A. Bachofen, Basel,
milling, and high-shear liquid impaction methods, such Switzerland) (Jinno et al., 2006, 2008), and yet others
as high-pressure homogenization, are highly effective have been developed in-house based on modifying
in generating and stabilizing newly formed nano- existing milling equipment (Nekkanti et al., 2009;
crystals and are increasingly used to produce particles Juhnke et al., 2010; Niwa et al., 2011b; Plakkot et al.,
in the nanometer range (Van Eerdenbrugh et al., 2008; 2011). These new methods have not yet been used to
Merisko-Liversidge and Liversidge, 2011). In addition, support the manufacture of a marketed product.
several bottom-up approaches to nanoparticle genera- Early descriptions of pearl milling suggested that
tion have been described and are finding increasing ap- long milling times (several days) may be required to
plication. Nanosized particles are produced in bottom-up produce nanosized particles (Merisko-Liversidge et al.,
approaches by controlling the rate of drug precipitation 1996; Muller et al., 2001; Hu et al., 2004a). However, it
from supersaturated solutions (Chan and Kwok, 2011). is apparent from the more recent literature that nano-
1. “Top-Down” Particle Size Reduction. scale particles may be achieved for some compounds
a. Pearl Milling. Pearl milling (alternatively de- after milling for much shorter periods of minutes or
scribed as nanomilling or media milling) is perhaps the hours (Fig. 43) (Merisko-Liversidge and Liversidge,
most established technique for the production of drug 2011). Longer milling times are likely required only
nanocrystals and has been described in detail in when low milling speeds are used or during the milling
several recently published reviews (Rasenack and of crystals that do not easily fragment (i.e., very hard
Muller, 2004; Peltonen and Hirvonen, 2010; Merisko- or very soft crystals). One concern surrounding the use
Liversidge and Liversidge, 2011). of pearl milling is the potential erosion of the milling
In brief, pearl milling involves the continuous media leading to contamination of the final product.
stirring and wet milling of an aqueous drug slurry The milling media must therefore be shown to with-
with specialized—and extremely hard and durable— stand the milling process to avoid potential erosion and
milling “pearls.” Nanocrystals smaller than 400 nm in product contamination (Merisko-Liversidge et al., 2003).
diameter may be produced, and particle size reduction PolyMill, the milling medium used in the NanoCrystal
is achieved via a combination of high-energy shear and platform, is said to be entirely resistant to the milling
impaction forces between the milling pearls and the process (Merisko-Liversidge and Liversidge, 2011),
solid drug particles. The addition of a polymer or sur- although final formulations for parenteral delivery are
factant stabilizer(s) to the aqueous slurry is essential routinely assessed for possible impurities.
Strategies for Low Drug Solubility 407

Fig. 43. Particle size reduction during milling. The particle size reduction of naproxen is shown in the processing chamber of the media mill. With
increased milling time, the particle size distribution profile using laser light diffraction decreases dramatically (i.e., shifts to the left) and narrows. The
particle size distribution curves are as follows: (A) the initial particle size of naproxen before processing, (B) 15 min postprocessing, (C) 30 min
postprocessing, and (d) 60 min postprocessing in a media mill. The time required to process a poorly water-soluble compound is dependent on
properties of the molecule, properties of the chosen stabilizer systems, and various processing parameters. For naproxen stabilized with povidone,
a residence time of 60 min provides a nanosuspension with a mean particle size diameter (DMEAN) of 200 nm. Adapted from Merisko-Liversidge and
Liversidge (2011).

Most drug powders are applicable to dry- and wet- of polymorphic transitions (crystalline-crystalline or
milling techniques. However, those with low glass tran- crystalline-amorphous) during the milling process.
sition or melting temperatures (typically below 100°C) b. High-Pressure Homogenization. High-pressure
show a tendency to deform plastically under stress homogenization (HPH) is widely used to produce nano-
(rather than fracture) and are therefore at greater risk sized particles and was recently reviewed (Keck and
408 Williams et al.

Muller, 2006; Muller et al., 2011a). Two different HPH particle interactions and, therefore, improves particle
techniques may be used to produce nanoparticles via size reduction (Muller and Bohm, 1998). However, this
a top-down approach, namely, microfluidization and benefit diminishes at loadings .10% w/w drug as
piston-gap HPH; the latter is the most widely reported a result of difficulties in stabilizing highly concentrated
technique. suspensions (Rabinow, 2004) and the increased energy
Piston-gap HPH, involves the passage of a stabilized required to process highly viscous particle slurries
aqueous drug suspension through a cylinder of approx- (Krause and Muller, 2001).
imately 3 cm in diameter before high-velocity stream- Methods of HPH have been widely described for the
ing at high pressure into a narrow gap of around a 25- development of parenteral and oral formulations
mm diameter. The dramatic increase in fluid pressure containing poorly water-soluble drugs (Peters et al.,
(up to 15,000–30,000 psi) causes water to boil at room 2000; Krause and Muller, 2001; Kayser et al., 2003;
temperature (in accordance with Bernoulli’s principle), Moschwitzer et al., 2004; Hecq et al., 2005, 2006; Wang
and this leads to cavitation of fluid at the particle et al., 2010b; Li et al., 2011c; Sahoo et al., 2011; Sun
surface. These cavities subsequently collapse as the et al., 2011; Wang et al., 2011a) (see also Table 23).
suspension leaves the narrow vessel (and as the pres- HPH has been suggested to be particularly suitable for
sure drops below the saturated vapor pressure of the parenteral delivery as there is no risk of contamination
liquid), and the resultant shockwaves cause particle from milling media and the high-pressure environment
disintegration into smaller nanocrystals (Muller and is able to eliminate potential microbial contaminants
Bohm, 1998; Muller and Junghanns, 1998; Patravale (Krauel-Goellner, 2008).
et al., 2004; Keck and Muller, 2006). Disadvantages of the HPH technique include the
Proprietary technologies that use HPH include the need for prolonged processing times when multiple
DissoCubes platform (SkyePharma AG, Muttenz, homogenization cycles are required to reach the target
Switzerland) and NanoPure (PharmaSol, Berlin, Ger- particle size (Gao et al., 2010b; Lai et al., 2009; Muller
many). The latter technology also permits the use of and Junghanns, 1998) and higher minimum batch
nonaqueous media, thereby allowing for the generation sizes compared with pearl milling by virtue of the
of water-free nanoparticle suspensions that may be higher void volume of most piston-gap homogenizers
filled into capsules. A number of piston-gap homoge- (Krauel-Goellner, 2008). The drug must also be pre-
nizers have also been described (Patravale et al., homogenized (at low pressure) or milled before HPH to
2004) and include the APV Micron Lab 40 (APV prevent blockage of the aperture in the piston gap
Deutschland GmbH, Lubeck, Germany), the EmulsiFlex- (Sahoo et al., 2011). Alternatively, drug crystals with
C3 and C5 (Avestin Inc., Ottawa, ON, Canada), and the small particle diameter may be harvested from
Stansted 7400 (Stansted Fluid Power Ltd., Stansted, supersaturated solutions (the basis for the bottom-up
UK). Piston-gap HPH is readily scalable, and similar approaches to nanoparticle production are described in
homogenizers are widely used to manufacture phar- more detail in the section below) and further processed
maceutical and food-grade emulsions (e.g., milk) at by HPH. This combination bottom-up/top-down ap-
commercial scale. proach is central to the NanoEdge technology (Baxter
The particle size of the nanosuspensions formed by International Inc, Deerfield, IL), which generates thin
HPH typically ranges from 40 to 500 nm (Muller et al., and fragile needle-like crystals (by precipitation) that
1999; Muller et al., 2001), although several process are highly susceptible to the subsequent homogeniza-
variables affect the size of the final product. First, tion process (Fig. 44) (Kipp et al., 2003; Rabinow, 2004;
owing to the high velocity of material transfer during Rabinow et al., 2007).
HPH, the residence time within the piston gap is short. 2. “Bottom-up” Nanoparticle Assembly. Bottom-up strat-
Therefore, material must be repeatedly cycled through egies for microparticle and nanoparticle production are
the homogenizer (Moschwitzer et al., 2004; Sun et al., based on the controlled precipitation or crystallization
2011) until the target particle size is attained or until of drug from a supersaturated solution (D’Addio and
no further change in size is evident with repeated Prud’homme, 2011). The first step in this process is the
cycling (Sahoo et al., 2011). Typically, 10 to 20 homo- generation of a supersaturated solution, followed by the
genization cycles are sufficient to achieve the minimum formation of crystal nuclei, which subsequently act as
particle size (Xiong et al., 2008; Gao et al., 2010a; a focal point for crystal growth. Crystal growth and
Pardeike and Muller, 2010), although larger cycle phase separation of solid drug subsequently continue
numbers may be required for drug crystals that until drug concentrations in solution reach the satu-
fragment poorly (Moschwitzer et al., 2004; Gao et al., rated solubility.
2010b). Increased pressure may be used to reduce the Attainment of a high degree of supersaturation is
number of homogenization cycles, however, this incurs critical to the success of nanoparticle formation as it
greater cost and potential wear to the apparatus encourages the formation of many small nuclei, the
(Mishra et al., 2009). The use of more concentrated ongoing growth of which is restricted by rapid de-
suspensions also provides a greater number of particle- pletion of the supersaturated drug pool (through
Strategies for Low Drug Solubility 409

Fig. 44. Engineering breakable crystals with a combination of microprecipitation and homogenization. Crystal morphology of raw drug material is
modified to facilitate breakage into smaller nanoparticles. (A) Crystals of starting raw material are too large and hard to run efficiently through
a homogenizer. (B) The raw material is solubilized, filter sterilized, and precipitated to yield crystals of needle-like morphology, which are easily broken
during homogenization. (C) Homogenization yields nanoparticles suitable for parenteral injections. Reprinted by permission from Macmillan
Publishers Ltd: Nat Rev Drug Discov (3:785–796), copyright (2004).

extensive nucleation) (Hu et al., 2004a; Chan and possible but may be eliminated by subsequent spray-
Kwok, 2011; D’Addio and Prud’homme, 2011; de Waard drying; however, this may also (unintentionally)
et al., 2011). In contrast, at lower degrees of supersat- yield amorphous drug particles (Rasenack and
uration, the generation of smaller numbers of crystal Muller, 2002).
nuclei and slower depletion of the supersaturated pool Supercritical fluids (SCFs) may be used as an
allow crystal growth to continue for longer periods, in alternative to organic solvents in a number of phar-
turn generating larger particles. Distinction between maceutical applications, including controlled crystalli-
controlled precipitation/crystallization methods is depen- zation. SCFs are typically gases at room temperature
dent on differences in the way in which supersaturation and pressure, but they exist at temperature and
is achieved (e.g., solvent shift or solvent evaporation) and pressures above a critical point in a phase that is
the nature of the solvents used. A number of the most neither gaseous nor liquid but shares properties of
common approaches to the bottom-up production of both. SCFs therefore have the solvent properties of
microparticulate and nanoparticulate systems are de- liquids but behave in many other respects like gases
scribed as follows. (Palakodaty and York, 1999; Pasquali et al., 2006).
a. Controlled Crystallization Via Solvent Shift. SCFs are of particular interest to particle synthesis
In solvent-shift methods, solvent containing drug at methods as it is possible to fine-tune their solvent
high concentration is mixed with an antisolvent (with properties via small changes in pressure and temper-
limited solvency for the drug), causing the solubility of ature (Phillips and Stella, 1993). The gaseous proper-
the drug in the mixed solvent to drop. This leads to ties of SCFs also afford rapid and efficient mixing
supersaturation and ultimately drug crystallization. (Chan and Kwok, 2011). Carbon dioxide (CO2) is the
The selection of a solvent-antisolvent combination that most widely used SCF for pharmaceutical applications
generates a high degree of supersaturation is critical to because of its moderate (and therefore practical)
the success of the solvent-shift method. The solvents critical temperature (304.2 K, 31.1°C) and pressure
should also be readily miscible as slow mixing will (7.38 MPa; 1070 psi). Supercritical CO2 is also non-
prevent the rapid establishment of a uniform degree of toxic, nonflammable, inexpensive, and can be recycled
supersaturation in the solvent mixture. The latter is during supercritical techniques (Bahrami and Ranj-
essential for the production of small particles with barian, 2007).
narrow particle size distribution. SCFs have been used for a number of years as
The Hydrosol solvent-shift technique was first a means of microparticle isolation via the rapid ex-
proposed by List and Sucker (1988) and involves pansion of supercritical solution (RESS) process (Phil-
the addition of a drug-containing organic solvent to lips and Stella, 1993; Rasenack and Muller, 2004;
a continuously stirring aqueous solution (Rasenack Bahrami and Ranjbarian, 2007). In RESS, a drug-
and Muller, 2002, 2004; Matteucci et al., 2007). It containing SCF is heated and passed through a nozzle
represents a relatively simple technique for pro- into a precipitation chamber. Passage through the
ducing nanoparticles, although wide particle size nozzle results in a drop in pressure, supersaturation,
distributions are evident where mixing is nonuni- and ultimately drug crystallization (Phillips and
form. Solvent carryover in the recovered product is Stella, 1993; Pasquali et al., 2006). Although RESS is
410 Williams et al.

a viable approach for particle production, both at the antisolvent. Discussion of these factors is beyond the
laboratory scale and for large-scale manufacture scope of the current article, however, details are
(Subramaniam et al., 1997), particles are typically greater available in the following references: Reverchon et al.
than 1 mm in size, largely because of unrestricted par- (2002, 2007), Muhrer et al. (2006), and Sathigari et al.
ticle aggregation in the precipitation chamber (Moribe (2011).
et al., 2005; Shinozaki et al., 2006; Chingunpitak et al., b. Precipitation after Solvent Evaporation. Solvent
2008; Su et al., 2009; Chen et al., 2010). Attempts to evaporation provides an alternative means of gener-
overcome this problem to generate nanosized material ating a supersaturated drug solution during bottom-
have included incorporating a cosolvent into the SCF up particle generation methods. Rapid attainment of
(Thakur and Gupta, 2006) and submerging the output a high degree of supersaturation is again the main
nozzle into an aqueous surfactant solution (Young objective; therefore, a high surface area of the drug
et al., 2000; Turk et al., 2002). However, RESS remains solution must be achieved to allow for rapid evapora-
limited by the often poor solvency of SCFs for poorly tion of the solvent. This has conventionally been
water-soluble drugs. Combining the SCF with a co- achieved using spray-drying (Vehring, 2008) but un-
solvent may, in some cases, overcome this problem til recently has been limited to the production of
(Larson and King, 1986; Dobbs et al., 1987; Thakur and micron-sized particles. More recently, new-generation
Gupta, 2006); however, it is now increasingly common spray-dryers (e.g., the Buchi B90; BUCHI Labortechnik
to use SCFs as antisolvents in more traditional solvent AG, Flawil, Switzerland) permit the production and
shift methods of nanoparticle production (described as collection of particles in the range of 300 to 5000 nm
below). and require only milligram quantities of drug (Heng
Techniques that use SCFs as the antisolvent are et al., 2011). A common drawback of spray-drying,
broadly termed SCF antisolvent techniques (SAS). The however, is the potential for unintentional genera-
various SAS approaches differ in the way in which the tion of amorphous drug particles rather than nano-
SCF interacts with the drug-containing solvent (Jung crystals (Hu et al., 2004b; Rasenack and Muller,
and Perrut, 2001). For example, in the gaseous 2004), although for some applications, this may be
antisolvent (GAS) method, the SCF is passed through advantageous.
the drug solution as a gas (Muhrer et al., 2006; Other common variants of the solvent evaporation
Pasquali et al., 2006), whereas the aerosol solvent method include evaporative precipitation into aqueous
extraction system (ASES) or precipitation from com- solution (EPAS) where a heated drug solution (nor-
pressed antisolvent (PCA) method requires the drug mally an organic solvent of low boiling point to avoid
solvent solution to be sprayed as a fine mist into a cham- thermal degradation of the drug) is sprayed as fine
ber of compressed SCF. In PCA, the higher surface droplets into a heated aqueous solution containing
area of the aerosolized drug solution permits a faster various stabilizers (Hu et al., 2004a). The aqueous
rate of mass transfer into the antisolvent, leading to suspension obtained is then freeze-dried or spray-dried
rapid attainment of higher levels of supersaturation to generate solvent (water) free particles. EPAS has
than are possible with GAS techniques (Fusaro et al., shown recent utility in generating amorphous nanosize
2005). Indeed, recognition of the importance of the particles of cyclosporine (Chen et al., 2002), quercetin
speed of mass transfer (to generate high and uniform (Gao et al., 2011b), and itraconazole (Chen et al.,
supersaturation) led York and coworkers to develop the 2004a). A variety of other solvent evaporation techni-
solution enhanced dispersion by SCF (SEDS) method, ques for producing nanoparticles have been developed
where a two-channelled coaxial nozzle design permits and include electrospraying (Jaworek, 2007), cryogenic
a high-velocity stream of SCF (CO2) to disperse and spray-freezing into liquid (SFL) (Hu et al., 2004a), and
combine with a continuously flowing stream of drug- controlled crystallization during freeze-drying (CDFD)
containing solvent. In the SEDS method, conditions (de Waard et al., 2008), although these appear to have
within the mixing chamber are controlled such that the been less widely applied compared with the more
CO2 remains in its supercritical state and the fast rate established bottom-up techniques.
of mass transfer between the SCF and drug-containing Interestingly, there have been few published com-
solvent leads to rapid nucleation and subsequent parisons of top-down versus bottom-up approaches to
particle formation (Hanna and York, 1998; Palakodaty microparticle and nanoparticle production, although
and York, 1999; Pasquali et al., 2006). the greater commercial success that has been enjoyed
Processing variables that influence the character- by top-down methods may be attributable to more
istics of the harvested particles (e.g., shape, size, and facile scale-up (de Waard et al., 2011). Nevertheless,
size distribution) include those that affect the rate of owing to their simplicity, speed, and low cost, bottom-
mass transfer of the saturated solution, such as the up techniques are highly appropriate for laboratory-
drug solubility in the solvent, the pressure and scale and proof-of-concept experiments and, with wider
temperature within the precipitation chamber, and use, may emerge as viable large-scale production
the rate at which the drug solution mixes with the techniques.
Strategies for Low Drug Solubility 411

C. Oral and Parental Delivery of Formulations 47). A recent example of the use of nanosuspensions to
Containing Nanosized Drug Particles support dose escalation studies of an (unnamed) BCS
1. Oral Delivery. Examples of drug nanoparticle- class II model compound has also shown evidence of
based formulations used in the oral delivery of poorly significantly higher drug exposure from a nanosuspen-
water-soluble drugs are provided in Table 24 and sion compared with micronized formulations and
include the marketed products Rapamune (containing sufficient exposure to allow saturation of elimination
rapamycin) and Emend (containing aprepitant). Data at the highest dose (Sigfridsson et al., 2011b). Finally,
comparing the oral bioavailability of aprepitant after oral administration of 100 mg/kg nanosuspension of
administration to beagle dogs as a nanosuspension the very poorly water-soluble antiprotozoal, atova-
quone, to mice infected with Toxoplasma gondii,
(120 nm) and a conventional microsuspension (5.5 mm)
provided efficacy that exceeded that of a microsuspen-
are reproduced in Fig. 45 (Wu et al., 2004). In this
sion and was comparable to that of a 10 mg/kg i.v. dose
example, tablets containing micronized aprepitant had
(Scholer et al., 2001).
previously shown evidence of a positive food effect in
In contrast, a small number of studies have reported
human subjects (unpublished), and a food effect was
a reduction in drug exposure after administration of
also evident (3.2-fold increase in bioavailability) on
oral nanosuspensions compared with conventional
dosing the micronized drug as a suspension to fed dogs.
solubilized formations (Kesisoglou et al., 2007; Chiang
Administration of the NanoCrystal formulation, how-
et al., 2009), and in common with most delivery strat-
ever, resulted in a 4.3-fold increase in bioavailability
egies designed to enhance oral exposure, apocryphal
over the micronized drug in the fasted state and no
evidence suggests there are examples where nano-
difference in drug exposure after coadministration
suspensions fail to provide significant increases in
with food (Fig. 45). A lipid formulation showed similar
exposure, but they are not reported in the literature.
performance (Wu et al., 2004), but low solubility in
Nonetheless, liquid nanosuspensions are effective in
traditional lipidic excipients dictated that multiple many situations, are relatively simple, and allow the
capsules (.10) would have been required to achieve administration of high doses, with low quantities of
the projected human dose. excipient, and therefore, should be considered partic-
The microparticulate and nanoparticulate systems ularly during preclinical studies where required doses
previously shown to enhance itraconazole dissolution are likely to be high. Furthermore, satisfactory results
in vitro in Fig. 42 were also subsequently examined in after the preclinical use of a nanosuspension provides
vivo after oral administration to rats, and the differ- a relative facile transition point for the progression and
ences in dissolution were shown to translate to differ- scale-up of the formulations into a solidified nano-
ences in bioavailability (Fig. 46). Thus micronization suspension for clinical use.
(to 5.5 mm) resulted in a 6-fold increase in oral bioavail- 2. Nanosuspensions for Parenteral Delivery of Poorly
ability compared with a nonmicronized product, and Water-Soluble Drugs. The risk of embolism after
further particle size reduction to yield nanoparticles administration of particulate formulations has tradi-
with diameters of 350 or 700 nm (by high-pressure tionally precluded the use of suspensions as i.v.
homogenization) elicited further (6.8-fold) increases in formulations for poorly water-soluble drugs. Nano-
exposure compared with the micronized product (.40- suspensions, however, provide the potential for rapid
fold compared with nonmicronized drug) (Sun et al., in vivo dissolution and significantly reduced risks of
2011). In this case, the additional benefits of a reduc- embolism via the use of particles with diameters that
tion in particle size from 700 to 350 nm were relatively are much smaller than that of most capillaries (Wong
small. et al., 2008). As such, nanosuspensions are increasingly
A significant advantage of drug nanosuspensions is used during preclinical testing (Kipp, 2004; Rabinow,
the ability to incorporate readily into tablet and 2004; Kesisoglou et al., 2007; Rabinow et al., 2007;
capsule formulations and to use the same technology Sigfridsson et al., 2007), although progression into
to generate simple oral suspension formulations for human clinical assessment has been slow. At this point,
preclinical pharmacokinetic, efficacy, and toxicity stud- the only marketed example is a sustained release i.m.
ies (Merisko-Liversidge et al., 1996; Langguth et al., injection of paliperidone palmitate (Invega Sustenna or
2005; Merisko-Liversidge and Liversidge, 2008). For Xepilon;EU Janssen Pharmaceuticals); however, sev-
example, formulation of a high permeability and low eral i.v. nanoparticle formulations are currently under
solubility (,1 mg/ml) developmental compound in an development, for example, thymectacin (Theralux,
oral nanosuspension allowed the administration of Celmed), which is currently in phase I/II clinical trials
higher doses than were possible using a nonaqueous for the treatment non-Hodgkin lymphoma.
lipid formulation (750 versus 350 mg/kg) and sub- Interest in the use of parenteral nanosuspensions
sequently led to significantly increased drug exposure in stems from their potential to deliver high drug loads in
rats compared with both nonmicronized material and a rapidly dissolving form without the need for large
a lipid-surfactant mixture (Kesisoglou et al., 2007) (Fig. quantities of excipients that may be irritant or toxic
412 Williams et al.

TABLE 24
Selected particle size manipulation references describing effect on in vitro dissolution and in vivo performance
Median or Mean Particle Size Form Used during In Dissolution/Bioavailability Enhancement
and Stabilizer/Dispersants Used Vitro/in vivo Testing

Top-Down Via Pearl-Milling


Cilostazol (Jinno et al., 2006) 220 nm Suspension 6.2- to 6.6-fold increase in exposure
HPC and docusate sodium (AUC) in beagle dogs compared
with jet-milled and hammer-milled drug
Reduced food effect
Cilostazol (Jinno et al., 2008) 260 nm Wet-milled tablet 12.8-fold increase in bioavailability (5–64%)
HPC and docusate sodium in fasted beagle dogs over a commercial
D-Mannitol tablet
Diminished food effect
In vitro/in vivo correlation reported
Danazol (Liversidge and 169 nm Suspension 15.1-fold increase in AUC in beagle dogs
Cundy, 1995) PVP K-15 over a normal suspension
Phenytoin (Niwa et al., 2011b) 292 nm Spray-dried .95% drug dissolved within 10 min at
PVP K-30 and SLS suspension pH 1.2 and 6.8
D-Mannitol D-Mannitol aided redispersion of the
dried nanoparticles
Undisclosed model compound 280 nm Suspension High-dose nanosuspensions showed higher
(Sigfridsson et al., 2011b) HPMC 1500 cPs and exposure (1.65- to 2.1-fold increase in AUC)
PVP K-30, SDS in rats over a microsuspension formulation
Ibuprofen (Plakkot et al., 2011) 270 nm Suspension Both suspension and spray-dried suspensions
HPMC (6 cPs), PVP K-30 and spray-dried showed .85% release within 2 min in acidic
and SLS suspension media (pH 2)
Indomethacin (Liu et al., 2011) 394 nm/393 nm/986 nm Freeze-dried Nanosuspensions containing either stabilizer
Pluronic F68/Pluronic nanosuspension showed rapid dissolution (pH 5) despite
F127/Tween 80 significant differences in particle size.
Ziprasidone (Thombre et al., Particle size not described Suspension Nanosuspensions dosed to human subjects
2011) Pluronic F108/Tween 80 showed a diminished food effect compared
and Soybean lecithin with ziprasidone HCl (Geodon) (;1.3-fold
and ;2-fold increase in AUC respectively).
However, there was an increase in Cmax and
decrease in Tmax with the nanosuspension,
hence the nanoformulation and the branded
product were not bioequivalent.
Top-down via high pressure
homogenization
Itraconazole (Sun et al., 2011) 316 nm Freeze-dried Nanosuspension showed a 51-fold/7.8-fold
Pluronic F127 nanosuspension increase in AUC in rats compared with raw
and SLS powder and microsuspension, respectively.
See Fig. 42 and Fig. 46.
Deacety mycoepoxydiene 515 nm Freeze-dried Near complete (.80%) drug dissolution within
(Wang et al., 2011a) Lecithin, poloxamer nanosuspension 10 min across range of pH (1.2–7.0)
188, HPMC and PVP. compressed into
D-Mannitol tablets
Diclofenac (Lai et al., 2009) 580 nm Freeze-dried Improved dissolution rate over free drug,
Pluronic F68 nanosuspension drug:stabilizer physical mixtures and coarse
in gelatin capsules suspensions in water, and in simulated
gastric and intestinal fluids
Rutin (Mauludin et al., 2009) 727 nm Freeze-dried 80% release in water within 30 min
SDS nanosuspension Superior dissolution compared with a marketed
compressed into formulation and microparticle formulation
tablets.
Cefpodoxime proxetil 245 nm Spray-dried 1.6-fold increase in AUC in rabbits compared
(Gao et al., 2010b) Poloxamer 188, HPMC nanosuspension with a marketed oral suspension
E3-LV and glycerol 1.69-fold increase in Cmax
Spironolactone (Langguth 400 nm Suspension 3.3-fold increase in AUC of canrenone
et al., 2005) Dioctyl sulfosuccinate (spironolactone metabolite) over a
microsuspension formulation
However, higher exposures were obtained
using a solid lipid nanoparticle formulation
and combining suspensions with a surfactant.
Revaprazan hydrochloride 562 nm Freeze-dried 1.36- to 1.45-fold increase in AUC in rats
(Li et al., 2011c) Poloxamer 188 nanosuspension compared with the coarse suspension and
microsuspension
Nanosuspension reduced Tmax from over 3 h
(for the suspension/microsuspension) to
under 1 h
Bottom-up
Itraconazole (Mou 249 nm (via precipitation Spray-dried 1.8-fold increase in AUC in fasted rats over
et al., 2011) from solution). suspension Sporanox pellets
HPMC E50 No change in AUC, Cmax or Tmax when
D-Mannitol administered in the fed state
(continued )
Strategies for Low Drug Solubility 413
TABLE 24—Continued
Median or Mean Particle Size Form Used during In Dissolution/Bioavailability Enhancement
and Stabilizer/Dispersants Used Vitro/in vivo Testing

Naproxen (Turk and 560–820 nm (depending on Harvested Nanoparticles obtained via the RESS approach
Bolten, 2010) RESS conditions). particles showed only a moderate (1.2-fold increase) in
300 nm (RESSAS using dissolution rate over unprocessed drug,
PVP solutions) attributed to the production of rod-shaped
7.8 mm (RESSAS using particles
Tween 80 solutions)
Itraconazole (Sathigari 940 nm (via SAS) Harvested Direct mixing of SAS drug particles with lactose
et al., 2011) 820 nm (via SAS followed particles led to faster dissolution (;90% dissolution at
by mixing drug with 30 min compared with , 50% in the absence
fast-flo lactose, 6% drug of excipient)
loading)
860 nm (via SAS in Pluronic
F127 solution and mixing
drug with fast-flo lactose,
50% drug loading)
Cmax, maximal drug concentration; SLS, sodium lauryl sulfate; SAS, supercritical antisolvent; RESS, rapid expansion of supercritical solution; RESSAS, rapid expansion of
supercritical solution into aqueous solution; Tmax, time of maximal concentration.

(Kipp, 2004; Rabinow, 2004). A further advantage is drug solubility (leading to Ostwald ripening) or by
the potential to retain the chemical stability advan- interacting with incorporated stabilizers (Jacobs et al.,
tages associated with drug in the solid state. Physical 2000). Where it is not possible to achieve adequate
stability concerns, however, are evident where the stability in liquid nanosuspensions, suspensions may
formulation is present as a predispersed suspension be freeze-dried and then reconstituted with water or
(via aggregation and particle size changes) and where saline before use (Muller and Bohm, 1998). This step,
there is potential for nanosuspensions to contain drug however, introduces additional risks since removal of
in anything other than the most stable crystal form the suspending medium (on drying) may lead to the
(via polymorphic transitions). formation of particle aggregates. If particles are not
Stability to particle aggregation and Ostwald ripen- readily dispersed during reconstitution of the dried
ing is enhanced by the addition of polymeric or suspension, the aggregates may cause irritation during
surfactant-based stabilizer(s). Those that have been injection and reduced drug efficacy following a slower
used in parenteral nanosuspensions include PVP, rate of dissolution. Redispersion may be facilitated by
Pluronic F68, Tweens, and lecithin and are discussed the addition of a dispersing aid such as mannitol or
in greater detail in the following references: Kesisoglou sucrose before the drying process (Hecq et al., 2005;
et al. (2007) and Wu et al. (2011). Tonicity modifiers Van Eerdenbrugh et al., 2008).
(e.g., glycerol and sodium chloride) are also required in The preclinical use of nanosuspensions has been
most i.v. formulations and have the potential to pro- reported to yield little, if any, irritation or pain at the
mote instability in a nanosuspension either by altering site of administration (Xiong et al., 2008; Wang et al.,

Fig. 45. Comparison of plasma concentrations of MK-0869 (aprepitant)


in beagle dogs after administration of 2 mg/kg drug via a conventional Fig. 46. Plasma concentration–time profiles of itraconazole after oral
suspension and a NanoCrystal formulation in fasted and fed animals. administration of suspensions with various particle sizes at a dose of 30
Values are expressed as means (n 5 5) 6 SEM. Adapted from Wu et al. mg/kg body weight in rats. Values are expressed as means (n 5 3).
(2004). Adapted from Sun et al. (2011)
414 Williams et al.

2010b; Gao et al., 2011c), and in some cases, nano- to dissolve sufficiently rapidly in situ that the phar-
suspensions are better tolerated than cosolvent for- macokinetic profiles of a formulated drug are indis-
mulations (Gao et al., 2011c; Brazeau et al., 2011). tinguishable from administration in a cosolvent or
Further evidence for the favorable safety profile of cyclodextrin formulation. Indeed, many studies have
nanosuspensions may be drawn from the development observed similar pharmacokinetics after administration
of several nanosuspension formulations for use in the of a nanosuspension formulation compared with other
clinic, although, as described already, these are yet to parenteral formulations (Fig. 48) (Clement et al., 1992;
reach the market. Viernstein and Stumpf, 1992; Gassmann et al., 1994;
To avoid phlebitis and potentially more serious sys- Sigfridsson et al., 2007). However, there is also a
temic side effects such as hypersensitivity and pulmo- significant body of evidence that has described differ-
nary embolism, nanosuspensions for i.v. use require ences in drug pharmacokinetics after i.v. administra-
particles with diameters of less than 300–400 nm tion of nanoparticle formulations. These are usually
(Jacobs et al., 2000; Muller et al., 2001; Li and Zhao, manifest as a marked reduction in Cmax and an in-
2007). A risk of local irritation remains, however, if crease in mean residence times after nanoparticle
nanoparticles aggregate on contact with blood (Ara- administration, presumably reflecting the incomplete
mwit et al., 2006), and the i.v. compatibility of a nano- initial dissolution and altered distribution and elimi-
suspension should be assessed in a manner analogous nation profiles of particulate drug as opposed to drug in
to that commonly used for buffered/cosolvent/micellar solution (White et al., 2003; Rabinow et al., 2007; Wang
formulations (see Sections IV, VI, and VII; and Yalkowsky et al., 2010b; Zakir et al., 2011; Gao et al., 2011c). For
et al., 1998). A number of studies have also linked the example, the apparent plasma half-life of an itracona-
i.v. use of particulate formulations with hemodynamic
zole-cyclodextrin solution formulation (Sporanox) is
effects, including a reduction in arterial blood pressure
approximately 5 h after i.v. administration to rats but
(Slack et al., 1981; Douglas et al., 1987; de Garavilla
increased to 15 h after administration of a crystalline
et al., 1998; Sigfridsson et al., 2011a). Reducing the
nanosuspension (Fig. 49) (Rabinow et al., 2007). The
rate of administration or use of histamine premedica-
nanosuspension was also associated with a reduction
tion may reduce the severity of these and other adverse
in Cmax (Fig. 49), although at an equivalent dose, it
effects associated with parenteral use of nanosuspen-
showed superior antifungal properties (Rabinow et al.,
sions (de Garavilla et al., 1996, 1998; Aramwit et al.,
2007). Perhaps unsurprisingly, effects on pharmacoki-
2006).
The reduced particle size of nanosuspension formula- netics are highly dependent on the particle size of the
tions provides the opportunity for parenteral formulations nanosuspension (Gao et al., 2008; Wang et al., 2010b),
and amorphous (rather than crystalline) nanosuspen-
sions appeared to show smaller reductions in Cmax

Fig. 47. Comparison of nanosized active pharmaceutical ingredient to


a nonaqueous vehicle suspension in support of toxicology formulation Fig. 48. The mean plasma levels of AZ68 versus time after i.v.
development for an insoluble compound (aqueous solubility ,1 mg/ml). administration of AZ68 in a PEG 400:DMA:water (1:1:1) solution, as an
Utilization of nanosizing enabled an increase in maximum feasible dose amorphous nanosuspension, and as a crystalline nanosuspension to rats.
and resulted in significantly higher exposure compared with a lipid- The administered dose was 5 mmol/kg in all cases. Values are expressed
surfactant mixture (Merck; data on file). Adapted from Kesisoglou et al. as means (n 5 3). PEG, polyethylene glycol; DMA, dimethylacetamide.
(2007). Adapted from Sigfridsson et al. (2007).
Strategies for Low Drug Solubility 415

into phagocytic cells of the mononuclear phagocytic


system (MPS) (Ganta et al., 2009; Wang et al., 2011b),
leading to accumulation in MPS organs, such as the
liver and spleen (Muller et al., 2011a). The uptake of
nanoparticles by the MPS may therefore limit rapid
drug exposure to the site of action, although gradual
drug release post-phagocytosis has the potential to
provide sustained release back into the systemic
circulation (Kipp, 2004; Patravale et al., 2004; Rabinow
et al., 2007). The use of nanoparticulate systems also
provides a means of directly targeting disease states
within the MPS. The potential use of modified nano-
suspension formulations to form the basis of targeted
or site-specific delivery systems in a manner analogous
to that of liposomes or nanoparticles is outside the
scope of the current article but has been addressed
elsewhere (Vonarbourg et al., 2006; Muller et al.,
Fig. 49. Plasma itraconazole concentration time profiles after i.v. 2011a).
injection of itraconzole nanosuspension and Sporanox to rats. Values
are expressed as means (n 5 3). Adapted from Rabinow et al. (2007).
D. Summary
Top-down (e.g., milling) or bottom-up (e.g., precipi-
compared with solution formulations. Together, these tation) techniques for decreasing particle size are
data suggest that it is the requirement for particle effective and widely used techniques to increase the
dissolution that leads to the decrease in Cmax (Rabinow rate of dissolution of poorly water-soluble drugs. The
et al., 2007) after parenteral administration of nano- mechanism for enhanced dissolution is largely via an
suspension formulations. increase in surface area with reduction in particle size.
Nonetheless, by virtue of a reduction in Cmax, nano- The advent of technologies that are able to generate
suspensions may also yield fewer drug-related toxic nanosized (rather than micronized) drug particles has
events for compounds where toxicity is correlated with resulted in significant increases in surface area and
maximum plasma concentrations and, therefore, pro- dissolution rate and also the potential for additional
vide the opportunity to use higher drug doses during advantage via increases in apparent drug solubility.
dose escalation (Sharma et al., 2011) and clinical Increases in solubility with particle size reduction may
studies (Peters et al., 2000; Scholer et al., 2001; stem from the introduction of imperfections into the
Margolis et al., 2007; Rabinow et al., 2007). In contrast, crystal structure during high-energy nanomilling or
in some examples where nanoparticles have led to from changes in particle curvature, although the latter
a reduced Cmax, a reduction in clearance and a corre- effect is evident only in particles with diameters
sponding increase in total exposure are also evident. smaller than 1 micron. Increases in solubility are
Under these circumstances, and where toxicity is more relatively modest, however, and changes to dissolution
highly correlated with total exposure (rather than peak rather than solubility remain the dominant advantage
exposure i.e., Cmax), nanosuspension administration of nanocrystals.
may lead to greater toxicity. The propensity for altered The rapid dissolution rate of nanosized suspensions,
pharmacokinetics after i.v. administration also likely the ability to generate particle sizes significantly
precludes the use of nanoparticle formulations as an smaller than blood capillary diameters, and the low
intravenous control during absolute bioavailability irritancy of drug nanoparticles have also led to the use
estimations since the requirement for identical sys- of nanocrystals by parenteral as well as oral routes of
temic pharmacokinetic parameters (clearance, volume administration. Drug nanocrystals therefore provide
of distribution) after i.v. and oral administration might a flexible approach to oral and parenteral administra-
not be met. tion and can be used both preclinically as a suspension
Where the parenteral use of nanosuspensions is and clinically after incorporation into more traditional
associated with prolonged drug exposures, there is also tablet or capsule formulations. Particle size reduction,
the potential for development of a controlled-release however, is not always effective at increasing exposure,
delivery platform (van’t Klooster et al., 2010), with the and reaggregation phenomena may limit increases in
advantage of controlling the period of sustained release dissolution rate. The need for relatively high shear to
simply by modifying the particle size (Rabinow, 2004; reduce particle sizes to submicron levels also runs the
Gao et al., 2008; Baert et al., 2009; Wang et al., 2010b). risk of inducing changes to crystal form and the
However, intact nanoparticles in the systemic circula- introduction of amorphous character. Finally, nanosiz-
tion are likely to be eliminated rapidly after uptake ing requires the use of specialized equipment and, until
416 Williams et al.

recently, has been limited to application under license slowly dissolving crystalline itraconazole (Mellaerts
because of intellectual property considerations. None- et al., 2008; Van Eerdenbrugh et al., 2009). Other
theless, the use of a nanocrystalline drug as a means to marketed formulations that use solid dispersion
enhance the delivery of poorly water-soluble drugs is formulations are listed in Table 25.
an established technology, and examples of commercial Despite their potential benefits, the number of
oral and injectable products are evident. commercially available solid dispersions is relatively
low. This reflects the thermodynamic instability of
X. Solid Dispersions drug in the amorphous form and the inevitable com-
plexities of assuring drug stability, the inability to
The prospect of using solid dispersions as a means of process many solid dispersion formulations using
enhancing dissolution and oral bioavailability for conventional manufacturing methods such as granula-
poorly water-soluble drugs was first reported by tion, compression, and coating, and a lack of analytical
Sekiguchi and Obi in 1961 (Sekiguchi and Obi, 1961) characterization tools that can demonstrate, with
and has been widely explored over the subsequent 50 high confidence, similarity among batches of materials
years (Hancock and Zografi, 1997; Serajuddin, 1999; sufficient to provide quality specifications for accep-
Leuner and Dressman, 2000; Newman et al., 2012). tance and rejection of the drug product. A classic
Solid dispersions increase drug dissolution via several example of the difficulty of progressing thermodynam-
mechanisms, including a reduction in effective particle ically unstable formulations is provided by the antivi-
size, improved wetting, enhanced solubilization, and ral ritonavir, a product that was initially marketed as
elimination of the impact of lattice energy via stabili- a capsule formulation (Norvir; Abbott Laboratories)
zation of drug in the more soluble amorphous state. but was later withdrawn after the identification of
The mechanism of dissolution enhancement is largely a more stable (and less soluble) crystalline polymorph
dictated by the structure of the solid dispersion formed, (Bauer et al., 2001).
but it is underpinned in many cases by the fundamen- Research in this area over the last 10 years, how-
tal differences in solubility and dissolution behavior of
ever, provides good evidence that the preparation of
a crystal and a glass of the same material (Grantschar-
solid dispersion formulations with both improved
ova and Gutzow, 1986).
dissolution and acceptable stability is possible. In
The term solid dispersion covers a range of different
particular, there is a growing understanding of the
(but related) formulations, all containing drug dis-
mechanisms by which high-molecular-weight polymers
persed within an inert carrier matrix (typically water-
such as HPMC (and derivatives), PVP, and the
soluble sugars, polymers, or surface active emulsifiers).
methacrylates are able to attenuate the drivers of
The principal difference between solid dispersion
drug crystallization (Yoshioka et al., 1995; Van den
subtypes is the physical form of the drug and the
Mooter et al., 1998; Matsumoto and Zografi, 1999) and,
carrier. Drug is either suspended in the carrier as
phase-separated crystalline or amorphous particles, or therefore, enhance physical stability. The mechanisms
it exists as a homogeneous molecular mixture of of amorphous stabilization include a reduction in
(amorphous) drug and carrier. The carrier can exist molecular mobility by increasing local viscosity in the
in the amorphous or crystalline form.
Given the often wide differences in crystalline and TABLE 25
Commercial pharmaceutical products that use solid
amorphous drug solubilities, different classes of solid dispersion technology
dispersion often result in significant differences in
Drug Carrier
dissolution rate. Solid dispersions that contain drug in
the amorphous form usually show much faster disso- Sporanox (itraconazole; Janssen Pharmaceuticals) HPMC
Intelence (etravirine; Tibotec) HPMC
lution rates than formulations containing crystalline Prograf (tacrolimus, Astellas Pharma US, Inc) HPMC
drug, leading to significant improvements in drug Nivadil (nilvadipine; Fujisawa Pharmaceutical HPMC
Co., Ltd)
absorption (Doherty and York, 1987; Law et al., 1992, Certican or Zortress (everolimus; Novartis HPMC
2004; Kennedy et al., 2008; Van Eerdenbrugh et al., Pharmaceuticals)
Isoptin-SRE (verapamil; Abbott Laboratories) HPC/HPMC
2009; Li et al., 2011b; Newman et al., 2012). Solid Zelboraf (vemurafenib; Roche) HPMCAS
dispersions containing drug in the amorphous form are Incivek (telaprevir; Vertex Pharmaceuticals) HPMCAS
therefore often preferred. A notable example of a mar- Gris-PEG (griseofulvin; Pedinol Pharmacal Inc) PEG
Cesamet (nabilone; Valeant Pharmaceuticals) PVP
keted amorphous solid dispersion is that of Sporanox Rezulin (troglitazone; Parke-Davis) PVP
capsules (Janssen Pharmaceuticals), an antifungal pro- Kaletra (lopinavir and ritonavir; Abbott PVPVA
Laboratories)
duct containing the poorly water-soluble drug itraco- Norvir tablets (ritonavir; Abbott Laboratories)a PVPVA
nazole molecularly dispersed in HPMC and coated onto
HPC, hydroxypropyl cellulose; HPMC, hydroxypropyl methylcellulose; HPMCAS,
inert sugar beads. Oral administration of the Sporanox HPMC acetate succinate.
a
solid dispersion formulation results in significantly en- The original Norvir capsule formulation was withdrawn from the market, but
was subsequently relaunched as a PVPVA melt-extruded tablet (Norvir Film-coated
hanced oral bioavailability compared with the more Tablets) and as a lipid formulation (Norvir SEC).
Strategies for Low Drug Solubility 417

solid dispersion matrix (Hancock and Zografi, 1997; Here we describe the various solid dispersion sub-
Shamblin et al., 2000; Van den Mooter et al., 2001; Six types and the mechanisms of drug release that un-
et al., 2004), the generation of intermolecular inter- derpin solid dispersion utility. Common methods of
actions between polymer and drug that also restrict manufacture are also briefly described. Finally, the
molecular movement (Taylor and Zografi, 1997; Khougaz manner in which polymers may be used to maintain
and Clas, 2000; Vasanthavada et al., 2005), and the the stability of the amorphous state in solid dispersions
inhibition of crystal nucleation (Konno and Taylor, over prolonged storage periods is discussed.
2006; Lindfors et al., 2008; Marsac et al., 2008; Caron
et al., 2010). A. Classification of Solid Dispersions
These processes all limit drug transformation from Solid dispersions can be classified according to the
the desirable (from a solubility perspective) but ther- molecular arrangement of the drug within the carrier
modynamically unstable amorphous form to the less matrix and the properties of the carrier. In simple
soluble but stable crystalline form. The effectiveness terms, drug may be 1) partially dissolved with excess
of crystallization inhibition is linked closely to the drug suspended in the crystalline state, 2) partially
crystallization tendencies of the drug, the properties of dissolved with excess drug suspended in the amorphous
the formulation (carrier), and the drug-to-carrier ratio state, or 3) completely dissolved (i.e., molecularly
since increasing drug load decreases the quantity of dispersed). The solid dispersion matrix may be either
carrier available to stabilize the drug in the high- crystalline or amorphous. As such, there are six major
energy form. There is also the added risk that high types of solid dispersions (although mixtures of amor-
drug loading will promote phase separation within the phous and crystalline particles are also possible). These
solid dispersion, in turn increasing the risk of instability are described in Table 26, and each arrangement is
(Marsac et al., 2009; Qian et al., 2010a). Nonformulation shown schematically in Fig. 50 (van Drooge et al., 2006).
factors that affect stability include the method of solid Solid dispersions containing amorphous drug particles
dispersion manufacture (widely used techniques include or molecularly dispersed drug are usually described as
hot-melt extrusion and spray- or freeze-drying) (Broman amorphous solid dispersions and those based on an
et al., 2001; DiNunzio et al., 2010b) and the presence of amorphous carrier as glasses. Where drug is present as
moisture and other plasticizers (Sethia and Squillante, a molecular dispersion, formulations are described as a
2004; Marsac et al., 2008; Rumondor et al., 2009b). solid solution (where the carrier is crystalline) or a glass
Extensive, rapid advances in understanding of the solution (where the carrier is amorphous). Depending on
fundamental physics of the glassy state (Johari, 2000; the drug loading in the solid dispersion relative to drug
Angell et al., 2003; Ngai et al., 2010) have also contrib- solubility in the carrier, solid or glass solutions may be
uted to a better understanding of the physical stability supersaturated (see Section X.G.5).
of amorphous systems, in particular, the importance of The different classes of solid dispersion are described
the kinetics of relaxation of high-energy solids (Lun- in the following sections, and systems where the
kenheimer et al., 2000; Prevosto et al., 2009). Increased carrier is crystalline (i.e., types 1–3 in Table 26) are
appreciation of the kinetic aspects of structural relaxa- discussed first. The simplest of these systems comprise
tion provides a platform to predict more accurately the crystalline drug suspended in a crystalline carrier and
molecular mobility of materials at temperatures below are commonly referred to as eutectic mixtures. Eutectic
the glass transition temperature (Bras et al., 2008; mixtures are combinations of crystalline components
Thayyil et al., 2008; Adrjanowicz et al., 2010b; Johari (drug and a carrier) that are miscible in the molten
et al., 2011; Kaminski et al., 2011) and, therefore, to liquid state but show little or no miscibility in the solid
understand the determinants of the long-term physical state (Sekiguchi and Obi, 1961; Chiou and Riegelman.,
stability of pharmaceutical formulations containing non- 1970). A eutectic system has a single chemical com-
crystalline solids. position that becomes a solid at a lower temperature
Finally, the fate of the drug after release from a solid than any other composition made with the same com-
dispersion is a key indicator of formulation perfor- ponents. This unique composition is known as the
mance, in particular, the propensity for drug recrys- eutectic composition and the temperature as the eutec-
tallization from the supersaturated solutions that are tic temperature. Below this eutectic temperature, a
generated by dissolution of amorphous drug. In the mixture of drug and carrier simultaneously crystallize
absence of stabilization, recrystallization occurs, lead- to form an intimate mixture of fine crystalline drug
ing to a decrease in the quantity of drug in solution particles within a crystalline carrier matrix. At all
available for absorption. However, the inclusion of other drug-carrier compositions (i.e., noneutectics), one
polymers within the formulation matrix has been component preferentially solidifies during cooling until
shown to reduce crystallization tendency post-dissolu- the remaining concentrations in the liquid phase reach
tion and to provide for significant increases in oral the eutectic composition, at which point the remaining
bioavailability in some cases (Tanno et al., 2004; drug and carrier (in the liquid phase) simultaneously
Friesen et al., 2008; Brouwers et al., 2009). crystallize.
418 Williams et al.

TABLE 26
Classification of solid dispersion subtypes according to the physical form of the drug and the carrier

Type Solid Dispersion Drug Carrier Total No. Physical Drug Stability
of Phases

1 Eutectic mixture Crystalline particles Crystalline 2 Stable


2 Solid amorphous Amorphous particles Crystalline 2 Unstable; crystallization risk
suspension
3 Solid solutiona Molecularly dispersed Crystalline 1 Stable (below crystalline solubility)
Unstable (above crystalline
solubility); crystallization risk
4 Glass suspension Crystalline particles Amorphous 2 Stable
5 Glass amorphous Amorphous particles Amorphous 2 Unstable; crystallization risk
suspension
6 Glass solution Molecularly dispersed Amorphous 1 Unstable; risk of phase-separation
and crystallization
a
May include continuous/discontinuous and substitutional/interstitial; see Leuner and Dressman (2000).

One of the first reports of the use of solid dispersions and Van den Mooter, 2009). Discontinuous solid sol-
as a formulation approach for poorly water-soluble utions are therefore the more relevant pharmaceutical
drugs described mixtures of sulfathiazole and urea system (Chiou and Riegelman., 1971). In regions of the
and classified the formulations as eutectic mixtures phase diagram where drug and carrier are miscible
(i.e., containing drug in the crystalline form) (Sekiguchi (i.e., where solid solutions are formed), drug is mixed
and Obi, 1961). Later work, however, suggests that within the carrier at a molecular level, and drug in this
sulfathiazole shows some degree of solid solubility in form exhibits the maximum attainable surface area for
urea, hence the original sulfathiazole-urea systems hydration (right hand panel; Fig. 50). The rate of drug
might more accurately be defined as a solid solution release from solid solutions is faster than that from
(i.e., type 3 in Table 26). In retrospect, it is likely that equivalent multiphase systems containing suspended
many of the solid dispersions that were first described solid drug particles, regardless of whether the sus-
as eutectic mixtures were either solid solutions, non- pended material is amorphous (e.g., type 2) or crystal-
eutectics, or monotectic systems, the latter comprising line (e.g., type 1).
combinations of drug and carrier in proportions where Solid dispersions containing drug dispersed in an
only the melting temperature of the drug is depressed amorphous carrier (i.e., types 4–6) may be classified
(i.e., where the melting temperature of the carrier is in a similar fashion to those where the carrier is
unchanged). crystalline, although in this case the solid dispersions
Within the class of solid dispersions described as are commonly referred to as “glasses” to reflect the non-
solid solutions (i.e., type 3), there are two subtypes: 1) crystalline character of the carrier. Thus, type 4 sys-
those in which the two materials are miscible at all tems, or glass suspensions, are analogous to eutectics
compositions in both the liquid and solid state (con- and contain crystalline drug suspended in an amor-
tinuous solid solutions) and 2) those in which complete phous carrier; type 5 systems, or amorphous glass
drug-carrier miscibility is observed only within certain suspensions, are related but contain amorphous drug
areas of the drug-carrier phase diagram (discontinuous particles suspended in an amorphous drug carrier; type
solid solutions) (Leuner and Dressman, 2000). Contin- 6 solid dispersions or glass solutions are analogous to
uous solid solutions are rare and are generally not solid solutions and contain a molecular dispersion of
reported in the pharmaceutical literature (Janssens drug in an amorphous carrier. The use of the term

Fig. 50. Schematic depicting the three different molecular arrangements of drug molecules within a solid dispersion. The panels illustrate the major
arrangement of drug in each solid dispersion subtype (Table 26). In types 1–3, the carrier is crystalline; in types 4–6, the carrier is amorphous. Note
that some molecularly dispersed drug will be present in each type (reflecting the solid solubility of the drug).
Strategies for Low Drug Solubility 419

glass to describe amorphous pharmaceutical systems the distinction between polymer stabilized nanoparticles
provides a facile point of differentiation between crystalline and nanoparticulate solid dispersions.
and noncrystalline carriers but is an oversimplification Finally, there is increasing interest in the generation
since glasses and amorphous solids, although both non- of amorphous drug formulations via drug adsorption to
crystalline, are thermodynamically distinct. Gupta (1996) a high surface area carrier. These systems stabilize
suggests a difference in amorphous and glassy materi- amorphous drug on the carrier surface and, where
als based on the degree of short-range order such that porous carriers are used, within the carrier pore struc-
short-range order in a glass is consistent with that in ture (Van Speybroeck et al., 2010b). Adsorbed formu-
a melt (allowing the material to undergo a glass lations such as these have occasionally been described
transition), whereas amorphous solids contain a higher as solid dispersions or surface solid dispersions (Kerc
level of short-range order precluding glass transition. et al., 1998; Watanabe et al., 2001; Takeuchi et al.,
In practice, however, both glasses and amorphous 2005; Wang et al., 2006). Differentiation between
solids effectively circumvent the limitations to solubil- traditional solid dispersions and adsorbed amorphous
ity provided by the crystal lattice, and a detailed systems can be made on the basis of the arrangement
discussion of differential thermodynamics is not war- of drug; in the case of solid dispersions, drug is molec-
ranted here. The interested reader is directed to the ularly dispersed or suspended throughout an inert
following references for more detail: Kauzmann (1948), carrier and crystallization is slowed primarily via an
Angell (1995a,b), and Gupta (1996). increase in viscosity and a decrease in drug mobility. In
1. Nontraditional Solid Dispersions. Several other adsorbed amorphous systems, drug crystallization is
formulation types have been described as solid dis- prevented through drug interaction with the surface
persions, including drug-excipient complexes, polymer- of the carrier (with the resulting spatial separation
stabilized nanoparticle formulations, and formulations preventing nucleation) and, in instances where the
containing amorphous drug where the amorphous form carrier is porous, the narrow dimensions of the fine
capillary network (that reduce mobility and limit crystal
is stabilized via adsorption to a high-surface-area, solid
growth) (Mellaerts et al., 2007). The use of high surface
carrier.
area adsorbents is described in more detail in Section
Solid drug-CD complexes may also contain amor-
XII.A.
phous drug, but in this case, drug molecules are
usually confined to the hydrophobic core of the CD B. Mechanisms by which Solid Dispersions Enhance
rather than being truly molecularly dispersed or sus- Dissolution Rate and Oral Bioavailability
pended within an inert matrix (Davis and Brewster,
1. Reduced Particle Size and Enhanced Drug Wetting
2004; Brewster and Loftsson, 2007; Stella and He,
and Solubilization. A common mechanism by which
2008). Drug-CD complexes that are further dispersed
all the solid dispersion subtypes in Table 26 enhance
within a polymer matrix better reflect the composition
dissolution is via an effective reduction in drug particle
of a solid dispersion, but they have the added size. Solid dispersions comprise drug (whether amor-
complexity of the drug-CD interaction. phous or crystalline) suspended or molecularly dis-
Polymer stabilized drug nanoparticles contain sim- persed in a carrier. In most cases, suspended drug
ilar components to solid dispersions and may also be particles have smaller particle sizes than are usual in
prepared using a similar approach, for example, solvent traditional solid dosage forms, and in the case of
evaporation techniques (Chan and Kwok, 2011). A a molecular dispersion, the particle size of drug in the
notable difference, however, is the lower drug-polymer carrier is minimized, theoretically, to the size of a single
ratio found in solid dispersion formulations (Rasenack molecule (Craig, 2002). The reduction in particle size
and Muller, 2002) where the polymer is required to provides for an increase in effective surface area for
disperse the drug molecularly (thus rendering it non- dissolution (in accordance with eq. 9). Secondary precip-
crystalline) and to assist in solubilization and wetting. itation from the supersaturated solutions that may be
In contrast, in most nanoparticle formulations, polymers formed by the dissolution of amorphous drug is also
are included to stabilize the dispersion and to decrease possible but commonly results in the formation of
agglomeration of small drug particles (Van Eerdenbrugh suspensions with relatively small particle size (Friesen
et al., 2008; Wu et al., 2011). Furthermore, the physical et al., 2008; Alonzo et al., 2011; Kanaujia et al., 2011).
form of the drug in a nanoparticle formulation is usually Indeed, it is often difficult to distinguish between an
crystalline, whereas in most recently developed solid increase in dissolution that stems from a reduction in
dispersions, drug is present in the amorphous form. The the particle size of the aggregates present in the original
fact that solid dispersions (where drug is embedded in solid dispersion and the size of particles that may be
a polymer matrix) may also be isolated in nanoparticle formed in situ via precipitation on dissolution. This is
or microparticle form, especially after spray-drying (Won further complicated by the realization that some level
et al., 2005; Kennedy et al., 2008; Wu et al., 2009; of particle aggregation is also likely (Simonelli et al.,
Dontireddy and Crean, 2011), adds further complexity to 1976; Alonzo et al., 2011). Nonetheless, in most cases,
420 Williams et al.

an increase in the dissolution rate resulting from a de- mixtures of nifedipine with PEG 4000 and 6000
crease in the size of dissolving drug particles is evident. provide rates of dissolution equivalent to that of a solid
The use of highly water-soluble carrier materials dispersion, even though the fusion of drug and polymer
also promotes rapid influx of hydration media into during solid dispersion manufacture leads to nifedipine
the solid dispersion matrix, resulting in an increase transformation to a metastable and more soluble
in wetting. To take advantage of this effect, 1st- physical form (Save and Venkitachalam, 1992). Zer-
generation solid dispersions used low-molecular- rouk et al. (2001) also report that although the
weight, highly water-soluble carriers such as urea bioavailability of carbamazepine is lower after admin-
(Goldberg et al., 1966a,c), aliphatic short-chain carbox- istration to rabbits in a PEG 6000 physical mix when
ylic acids such as citric and succinic acid (Goldberg compared with an equivalent solid dispersion, the dif-
et al., 1966b; Summers and Enever, 1976), and sugars ferences in exposure were not statistically significant.
such as sucrose, trehalose, sorbitol, and mannitol Nonetheless, and realizing that examples to the
(Allen et al., 1978; Arias et al., 1995; Okonogi et al., contrary are evident, in most cases, solid dispersion
1997; Valizadeh et al., 2004). In one of the earliest formulations outperform drug-polymer physical mix-
studies, Sekiguchi and Obi (1961) described an in- tures (McGinity et al., 1984; Dannenfelser et al., 1996;
crease in the rate of dissolution of sulfathiazole when Law et al., 2001; Jun et al., 2005; Han et al., 2011;
the drug was fused with urea and then crash-cooled to Kanaujia et al., 2011; Khan et al., 2011; Li et al.,
form a solid dispersion (when compared with a physical 2011b; Moes et al., 2011). Thus, no evidence of drug
mixture of drug and urea). The improvement in dis- dissolution over a 30-min period was reported for a phy-
solution rate was attributed to the rapid dissolution of sical mixture of itraconazole-HPMC, whereas a hot-
the eutectic and liberation of drug in the form of a melt extruded solid dispersion containing drug in the
microcrystalline dispersion and was ultimately shown amorphous form released ;80% drug within the same
to improve oral absorption in human subjects. period (Verreck et al., 2003b). Amorphous ritonavir-
A growing awareness of the potential benefits asso- PEG 8000 solid dispersions also show superior rates of
ciated with the use of polymeric or surfactant-based dissolution compared with rates of equivalent physical
solid dispersions and the relatively high melting tem- mixtures (Fig. 54A) (Law et al., 2004). In both these
peratures of many sugars (which complicates solid cases, it is likely that the advantage of the solid
dispersion manufacture via melting/fusion techniques) dispersion stems from differences in solubility result-
(Leuner and Dressman, 2000) have resulted in a gen- ing from isolation of drug in the amorphous form. The
eral decline in the use of 1st-generation carriers. advantages that accrue from the isolation of drug in
Second-generation polymeric carriers—including PVP, the amorphous form are described in more detail in
PEG, polymethacrylates, and HPMC (and its derivatives) Section X.B.2.
—and 3rd-generation carriers including self-emulsifying Third-generation solid dispersions include surface-
and surfactant-based excipients are now considerably active excipients that promote emulsification of the
more common (Serajuddin, 1999; Vasconcelos et al., formulation and solubilization of incorporated drug.
2007). Emulsifiers are widely used in solid dispersions either
Polyethylene glycol (PEG) was the first widely ex- alone or in combination with polymers and can ac-
plored 2nd-generation polymeric excipient and formed celerate drug release through enhanced drug wetting
the basis for the first commercial solid dispersion pro- and solubilization (Serajuddin et al., 1988, 1990).
duct (Gris-PEG, Pedinol Pharmacal Inc., Farmingdale, Emulsifiers also aid drug release by enhancing the
NY). PEG has a low melting temperature (the melting ability of the formulation to disperse adequately on
point of PEG 20 000 is between 60 and 63°C) and can hydration (Serajuddin, 1999). Some of the first studies
be readily fused with drug by quench-cooling a melt to explore solubilizing solid dispersions examined the
or via hot-melt extrusion (Craig, 1990; Leuner and utility of incorporation of polyethoxylated surfactants,
Dressman, 2000). PEG-based solid dispersions enhance including Labrasol, Tweens, and Gelucire, and showed
the dissolution of many low solubility compounds (Law improvements in dissolution rate and oral bioavail-
et al., 1992, 2001; Saers and Craig, 1992; Save and ability over nonsolubilizing solid dispersions (Serajuddin
Venkitachalam, 1992; Lloyd et al., 1997; Van den et al., 1988, 1990; Veiga et al., 1993; Sheen et al., 1995;
Mooter et al., 1998; Verheyen et al., 2002; Valizadeh Khoo et al., 2000).
et al., 2004; Baird and Taylor, 2011), however, benefits Solid emulsifiable glasses have also been prepared
over and above simple physical mixtures of drug and by drying emulsions (containing dissolved drug) in the
polymer are not always apparent. For example, PEG presence of sugars, with or without the inclusion of
6000–based solid dispersions increase the dissolution surfactants. These systems are closely related to solid
rate of diazepam and temazepam compared with dispersions (and may also be prepared by similar
crystalline drug, however, the rate of dissolution is techniques, for example, solvent evaporation); the dif-
no better than that of equivalent PEG-drug physical ference is the substitution of a dispersed oil phase
mixtures (Verheyen et al., 2002). Similarly, physical containing dissolved drug for the more typical
Strategies for Low Drug Solubility 421

dispersion of crystalline or amorphous drug (Myers and energy difference between amorphous and crystal-
Shively, 1993). Emulsifiable glasses self-emulsify on line drug and reduce the solubility advantage com-
hydration to form a fine drug-in-oil dispersion (Myers pared with theoretical predictions (Murdande et al.,
and Shively, 1992; Porter et al., 1996) and provide for 2010a,b).
equivalent drug exposure for cyclosporine after oral ad- Despite the potential for errors in the estimation of
ministration to dogs compared with traditional liquid solubility, significant solubility advantages remain for
self-emulsifying formulations (Porter et al., 1996). noncrystalline drug, and these are manifest through
A number of surface-active agents have been in- increases in dissolution rate and, ultimately, oral
corporated into solid dispersions, including ionic (e.g., bioavailability. This solubility advantage is the basis
sodium lauryl sulfate) and nonionic surfactants (e.g., for much of the utility of amorphous solid dispersions.
Pluronic block copolymers) (Chen et al., 2004b; Newa 3. Maintenance of Supersaturation after Drug
et al., 2007; Lakshman et al., 2008; Badens et al., Dissolution. Drug dissolution from dosage forms con-
2009), and these excipients offer additional benefits to taining high-energy polymorphs or amorphous solids
solid dispersion performance, including promotion of typically leads to the attainment of drug concentra-
drug mixing/miscibility within the solid dispersion tions in solution that are supersaturated with respect
(i.e., promotion of solid solution formation) (Dannenfelser to the saturated solubility of the stable crystal form.
et al., 1996; Wulff et al., 1996; Shin and Kim, 2003; The attainment of supersaturation is beneficial in
Linn et al., 2012). Since many of the surface active providing a greater concentration of dissolved drug and
agents used in the formation of solid dispersions of this in generating drug solutions with increased thermody-
type are liquid or semisolid at room temperature, manu- namic activity. The induction of supersaturation in
facture into traditional solid dosage forms (e.g., tablets) solution has been referred to as a concentration
is difficult unless high-molecular-weight polymers are “spring” and may occur under many conditions, in-
incorporated. For example, mixtures of Tween 80 and cluding dissolution of amorphous drug, but also the
PEG (e.g., PEG 1000, 1450, 3350, and 8000) have been dispersion or digestion of lipid-based dosage forms,
shown to form solid formulations up to 75% surfactant the dissolution of salts of weak acids or bases, and after
(Morris et al., 1992). Alternatively, direct filling of drug- the gastric emptying of basic drugs that are soluble in
carrier melts into hard gelatin capsules has also been acidic gastric fluid, but less soluble in the intestine
attempted (Serajuddin, 1999). (Fig. 51) (Guzmán et al., 2004; Kostewicz et al., 2004;
2. Administration of Drug in the Amorphous Form. Brouwers et al., 2009; Warren et al., 2010; Porter et al.,
Amorphous solids lack the ordered molecular lattice 2011; Shono et al., 2011; Anby et al., 2012).
and defined molecular configuration and packing of Supersaturated solutions provide an advantage with
crystalline solids of equivalent chemical composition. respect to drug absorption but are unstable and at risk
The strength of intermolecular solid-state interactions of precipitation. The rate of drug precipitation from
in an amorphous solid is therefore lower than in supersaturated solutions can be markedly slowed by
crystalline solids. From the solubility equations in the presence of materials (including polymers) that in-
Section I.B.1, a reduction in intermolecular forces hibit crystal nucleation or slow crystal growth (Vande-
in the solid state contributes to an increase in solubility. cruys et al., 2007; Brouwers et al., 2009; Warren et al.,
Indeed, such is the importance of solid-state interac- 2010). Consistent with the “spring” analogy for super-
tions to solubility that the solubility gain attained by saturation induction, the reduction in the rate of drug
the use of drug in the metastable amorphous form may precipitation from supersaturated solutions has been
be up to several orders of magnitude. However, differ- referred to as a “parachute” and the combined mech-
ences in experimentally determined amorphous and anism of supersaturation generation and precipitation
crystalline solubilities are in practice often much lower inhibition as a “spring and parachute” (Guzmán et al.,
(Hancock and Parks, 2000; Murdande et al., 2011a). 2004, 2007) (Fig. 51).
This has been suggested to reflect the complexities of The potential for the polymers that are used to
measuring the “equilibrium” solubility of a thermody- fabricate solid dispersions also to inhibit drug pre-
namically unstable amorphous material, such that cipitation from supersaturated solutions post dissolu-
spontaneous crystallization results in underestimation tion is illustrated in Fig. 52A. In this example, three
of the initial (kinetic) solubility properties of the tacrolimus solid dispersions were manufactured using
amorphous form (Hancock and Parks, 2000; Bhugra different carrier polymers. In each case, the formula-
and Pikal, 2008; Murdande et al., 2010a,b, 2011a). tions led to consistent (and rapid) initial rates of
Recent work also suggests that historical predictions of dissolution (Yamashita et al., 2003). However signifi-
amorphous drug solubility (using the difference in heat cant differences in the ability of the polymers to main-
capacity between crystalline and amorphous drug) may tain supersaturation were subsequently evident. HPMC
have been overestimated since amorphous solids are prevented recrystallization more effectively compared
often hygroscopic and invariably absorb water. The with PVP and PEG 6000, and HPMC-based solid dis-
resulting drug-water interactions decrease the free persions were later shown to provide for a 10-fold
422 Williams et al.

Fig. 51. Schematic illustrating a hypothetical dissolution profile of


a drug existing in a high-energy form (e.g., the amorphous form) with
respect to time. The solid black line depicts the fast dissolution of a high-
energy physical form, such as amorphous drug (the “spring”), the
attainment of drug concentrations in excess of the crystalline drug
solubility, and supersaturation. The dashed red line shows the rapid
decrease in dissolved drug concentration in response to nucleation and
crystal growth. The potential benefit of a “parachute” (e.g., a polymer
precipitation inhibitor) is shown by the dashed green line where periods
of supersaturation are more sustained, reflecting a slower rate of
nucleation/crystal growth. The scheme here describes the generation of
supersaturation via dissolution of amorphous drug from a solid dispersion
formulation but also applies to the dissolution of amorphous drug from
cyclodextrin formulations (see Section VII) or from adsorbed “surface”
solid dispersions (see Section XII.A). Supersaturation is also evident after
the dissolution of cocrystals (see Section V.B) or salts (see Section IV)
under certain conditions and the dispersion and digestion of lipid-based
formulations (see Section XI). In all cases, polymers have been explored
as a mechanism of supersaturation stabilization.

increase in Cmax and AUC in dogs compared with ad- Fig. 52. (A) Dissolution profiles of crystalline tacrolimus and tacrolimus
ministration of crystalline drug (Fig. 52B) (Yamashita solid dispersions containing HPMC, PVP, or PEG 6000. Each solid
dispersion formulation corresponding to 70 mg as tacrolimus was
et al., 2003). HPMC plays a similar role in enhancing dissolved in Japanese Pharmacopeia gastric fluid (pH 1.2), and the test
the bioavailability of the poorly water-soluble weak base carried out using the paddle method at 200 rpm (37 6 0.5°C). Values are
expressed as means (n 5 2). (B) The plasma concentration of tacrolimus
itraconazole after oral administration of the Sporanox in beagle dogs after oral administration of solid dispersions containing
formulation. In this case, drug dissolution under acidic HPMC compared with tacrolimus crystalline powders. Values are
conditions reflective of the gastric environment is un- expressed as means (n 5 6) 6 SEM. Each dosage form was administered
at the dose of 1 mg as tacrolimus. Adapted from Yamashita et al. (2003).
surprisingly rapid, but it is widely believed that the
robust bioavailability (;55%) of itraconazole after ad-
ministration of Sporanox is due, at least in part, to the for interactions with the aqueous medium. These
ability of HPMC to prevent drug precipitation as dis- properties assist in the maintenance of supersaturation
solved drug leaves the acidic stomach and enters the (Tanno et al., 2004). Figure 53 shows the ability of
less favorable solubilizing conditions (for a weak base) HPMCAS to maintain supersaturation for two experi-
in the intestine (Jung et al., 1999; Six et al., 2005; mental compounds at high (A) and low (B) drug loads
Augustijns and Brewster, 2012). (Curatolo et al., 2009).
Similar behavior with respect to the capacity for HPMCAS also has a high glass transition temper-
polymers to stabilize supersaturated drug has been ature (;120°C) and can therefore reduce mobility and
attributed to HPMCAS-based solid dispersions (Friesen protect against drug crystallization in the solid state
et al., 2008; Kennedy et al., 2008; Konno et al., 2008; [see Section X.F.3 and Section X.G]. In contrast, at
Curatolo et al., 2009; Murdande et al., 2011a). HPMCAS least for weak bases, the bioavailability enhance-
is unusual compared with many of the most commonly ments provided by solid dispersions of HPMCAS are
used polymers in oral formulations in that it shows unlikely to stem entirely from the maintenance of
some degree of amphiphilicity, with hydrophobic supersaturation since HPMCAS is largely insoluble
domains available for interaction with poorly water- in the acidic stomach contents, and on gastric empty-
soluble drugs and hydrophilic domains available ing, drug precipitation may occur before complete
Strategies for Low Drug Solubility 423

maximize the performance of the formulation (Augus-


tijns and Brewster, 2012). Support for this contention
is provided by the data described by Six et al. (2005),
where slower rates of itraconazole release from HPMC
solid dispersions led to better in vivo performance in
human subjects compared with solid dispersions pre-
pared using Eudragit E100 or mixtures of Eudragit
E100 and PVPVA 64 that showed faster in vitro rates
of dissolution. Solid dispersions of itraconazole pre-
pared using enteric polymers (Eudragit L100-55 and
Carbopol 947P) to reduce drug release in the stomach
also showed enhanced performance compared with
HPMC (Miller et al., 2008; DiNunzio et al., 2010a).
Specifically, the enteric polymers reduced the quantity
of the dose that was solubilized in the acidic gastric
content and, therefore, reduced the solubilized bolus
that was introduced into the higher pH environment of
the intestine on gastric emptying. This in turn lowered
the degree of supersaturation that was invoked by the
increase in pH in the small intestine and the drop in
itraconazole solubility (Miller et al., 2008; DiNunzio
et al., 2010a). Subsequent bioavailability studies using
an itraconazole-Eudragit L100-55 enteric solid disper-
sion and a conventional HPMC solid dispersion in rats
revealed that the enteric matrix resulted in prolonged
and improved exposure (although the results were
highly variable) (Miller et al., 2008). Similarly,
although the release of a sparingly soluble glycogen
phosphorylase inhibitor was faster from solid disper-
sions formed with cellulose acetate phthalate (CAP)
compared with HPMCAS, the highly supersaturated
Fig. 53. (A) Dissolution performance of solid dispersions containing
drug solutions that resulted from dissolution of the
“compound 2” and various polymers at 50% w/w drug loading. Dissolution CAP formulation could not be maintained, and the
was carried out using the syringe dissolution test in model-fasted duodenal slower releasing HPMCAS solid dispersion showed
fluid at 37°C with a 500-mg/ml total concentration (dissolved plus
undissolved drug). HPMCAS, hydroxypropyl methylcellulose acetate improved overall performance (Crew et al., 2007).
succinate; PVAP, polyvinyl acetate phthalate. (B) Dissolution performance b. Mechanisms by which Polymers Maintain Super-
of solid dispersions containing “compound 5” and various polymers at 10%
w/w drug loading. Dissolution was carried out using the microcentrifuge saturation in Solution. Precipitation from a supersat-
dissolution test in PBS at 37°C with a 200-mg/ml total concentration urated solution follows two critical steps: the formation
(dissolved plus undissolved drug). CAP, cellulose acetate phthalate; HPMC, of particle nuclei and subsequent particle growth
hydroxypropyl methylcellulose. Adapted from Curatolo et al. (2009).
(Macie and Grant, 1986; Gebauer et al., 2008; Lindfors
et al., 2008). Nucleation and particle/crystal growth in
dissolution of the HPMCAS matrix (Van Speybroeck solution is in large part controlled by the same drivers
et al., 2010b). of phase-separation or crystallization that occur in the
a. The Impact of Dissolution Rate and Supersatura- solid state. These include decreases in molecular
tion Ratio on Precipitation Inhibition. The drivers of mobility (mediated via changes in viscosity and molec-
drug precipitation from supersaturated solutions in- ular interactions between drug and carrier) and direct
clude the degree of supersaturation (where increased inhibition of nucleation and crystal growth (see Section
supersaturation increases the potential for nucleation X.F). However, differences in the nature of the matrix
and subsequent precipitation or crystallization) (Turn- in which precipitation or crystallization occurs (i.e., in
bull and Fisher, 1949) and the rate of dissolution (where solution or within the polymer) result in important
rapid attainment of highly supersaturated solution also differences in assessment and stabilization.
increases the chances of nucleation and precipitation). One example of a means of reducing the potential for
Formulation approaches that moderate the degree of precipitation from solution is via an increase in drug
supersaturation in solution by either reducing the su- solubility in solution as this will reduce the level of
persaturation ratio or reducing the rate of dissolution supersaturation and the thermodynamic drivers of
therefore increase the stability of the supersaturated drug precipitation. Supersaturation stabilization via
solution, reduce the likelihood of precipitation, and an increase in solubilization is largely applicable to
424 Williams et al.

surface-active carriers such as the surfactants and on the characteristics of a highly ordered molecular
emulsifiers that are used in 3rd-generation solid structure, phase separation of crystalline solid results.
dispersions. Polymeric carriers have also been sug- In contrast, where crystal structure is not maintained,
gested to enhance the stability of supersaturated drug is expected to precipitate in the amorphous form.
solutions via an increase in drug solubilization (Usui Whether assembly of amorphous or crystalline par-
et al., 1997), however, in most cases, (nonsurface ticles occurs by two completely separate paths or
active) polymers offer little, if any, direct enhancement whether an amorphous “preparticle” precedes the sub-
of equilibrium solubility (Konno et al., 2008; Warren sequent generation of a lower-energy drug crystal or
et al., 2010; Bevernage et al., 2011). Instead, polymers higher energy amorphous precipitate is unclear
reduce drug precipitation via slowing the kinetics of (Gebauer et al., 2008; Warren et al., 2010). Friesen
drug precipitation or crystallization (Rodriguez-Hornedo and coworkers, for example, speculate that most of the
and Murphy, 1999; Raghavan et al., 2001; Lindfors drug released from HPMCAS solid dispersions is
et al., 2008; Miller et al., 2008; Brouwers et al., 2009; stabilized in solution in the form of amorphous drug-
Curatolo et al., 2009; Warren et al., 2010; Kanaujia polymer colloidal assemblies that are approximately 20
et al., 2011). to 300 nm in diameter (Friesen et al., 2008). The
Increased solution viscosity has also been proposed presence of colloidal drug-polymer aggregates permits
as a mechanism by which polymers inhibit precipita- more sustained periods of supersaturation by decreas-
tion in both the solid-state and in solution (Miller et al., ing the collision frequency of monomeric, solvated free
2008; Gao et al., 2009). However, precipitation in- drug molecules, which in turn decreases the kinetics of
hibition in solution has been observed at polymer nucleation (Friesen et al., 2008; Gebauer and Coelfen,
concentrations that are too low to have significant 2011; Murdande et al., 2011b). Recent studies have
effects on solution viscosity (Usui et al., 1997; Warren described the formation of similar nanoparticulate
et al., 2010), and polymers with similar viscosity en- species in dissolution media after rapid drug release
hancement properties can show marked differences in from felodipine-HPMC (Alonzo et al., 2011) and
their ability to inhibit precipitation (Vandecruys et al., ketoconazole-PVP solid dispersions (Kanaujia et al.,
2007; Warren et al., 2010; Bevernage et al., 2011). 2011).
Taking into consideration the markedly higher molec- The spontaneous formation of self-assembled molec-
ular mobility of polymers in solution (compared with ular species in solution has been shown even for a
the solid state), it seems likely that polymer-mediated single solute and has been attributed to conditions of
precipitation inhibition in solution is mediated more sustained supersaturation (Ohgaki et al., 1992). These
commonly by direct effects on nucleation or crystal observations are consistent with a broadly applicable
growth rather than changes in viscosity. In this regard, concept of the formation of prenucleation clusters that
polymers may slow nucleation by interacting with drug have “molecular” character in aqueous solution and
in solution and by adsorption to the interface of the create multiple equilibria among the various species
emerging particle/crystal nucleus (Okimoto et al., 1997; present. This in turn decreases the collision frequen-
Suzuki and Sunada, 1998; Yamashita et al., 2003; Gao cies of the solvated, unimolecular species and results in
et al., 2004; Yokoi et al., 2005; Brewster et al., 2008a). the creation of sustained, supersaturated solutions
Polymer effects on nucleation and crystal growth in the (Gebauer and Coelfen, 2011).
solid state are described in more detail in Section X.F. Although the formation of structured polymer-drug
In situations where only particle growth is reduced, nanoassemblies in water appears to sustain the period
spontaneous nucleation leads to an initial decrease in of supersaturation, their presence complicates analysis
dissolved drug concentration (due to nuclei formation) since small aggregates are less likely to separate from
and a subsequent reduction in the rate of crystal free drug in solution using traditional centrifugal or
growth (Konno et al., 2008). The ability of a polymer to filtration-based separation techniques and may also
adsorb to a developing drug nuclei also has the retain the spectroscopic properties of the parent (Alonzo
potential to reduce crystal growth and to disrupt the et al., 2011; Van Eerdenbrugh et al., 2011). Whether
manner in which the crystal lattice is formed (Lindfors drug is completely dissolved in solution or dispersed as
et al., 2008). This may promote drug precipitation as nanosized aggregates is therefore difficult to define
a high-energy crystalline polymorph or in the amor- accurately (Murdande et al., 2011b). Furthermore,
phous form (rather than as the thermodynamically whether nanoaggregates reduce the thermodynamic
stable crystal). In both cases, the precipitated material activity of drug compared with a typical solution (and
is expected to show more rapid redissolution compared therefore reduce the driving force for absorption) is
with a precipitate containing the stable crystal. The also unknown. In a situation analogous to the solu-
physical form of drug as it phase-separates from bilization of drug in a microemulsion or micelle, the
supersaturated solution is therefore an area of growing aggregate is expected to provide a solubilized reservoir
interest (Friesen et al., 2008; Alonzo et al., 2011; Hsieh that is in rapid equilibrium with drug in free solution
et al., 2012; Ozaki et al., 2012). Where the solid takes (Friesen et al., 2008). However, there is also a risk that
Strategies for Low Drug Solubility 425

nanoaggregates may grow and ultimately promote subsequently removed, filtered, or centrifuged and dis-
drug crystallization. Judicious selection of the polymer solved drug concentrations determined by UV spectro-
is expected to reduce this possibility. photometry (Vandecruys et al., 2007; Curatolo et al.,
Interestingly, although classic nucleation theory 2009). Continuous monitoring of turbidity via nephe-
(CNT) is most commonly used to explain crystallization lometry can also be used to track precipitation in real
and provides a backdrop to the use of crystallization time (Warren et al., 2010). The receptor medium used
inhibitors in pharmaceutical formulations (see Section is usually water or a buffered solution, although more
X.G.2), it is increasingly clear that CNT fails to capture complex screening methods may be performed using
the complexity of nucleation in many cases (Kashchiev biorelevant media such as simulated gastric or intesti-
and van Rosmalen, 2003) and that alternative mech- nal fluids. Bevernage and coworkers recently compared
anisms may be more appropriate. These include models the ability of polymeric excipients to prevent pre-
where molecular aggregates are initially formed, fol- cipitation from supersaturated solutions using simple
lowed by the generation of postcritical nuclei and sub- buffers, simulated intestinal fluids (fasted/fed), and
sequent nucleation of the crystal phase (Gebauer et al., aspirated human intestinal fluid. In this case, the
2008; Meldrum and Sear, 2008) and two-step mecha- capacity of the polymer to reduce precipitation was
nisms where the initial formation of metastable clusters diminished when studies were run in aspirated in-
that are several hundred nanometers in size precedes testinal fluid (Bevernage et al., 2011).
crystal nucleation within the clusters (Vekilov, 2010). Of the carriers that are commonly used in the
The use of computational physics and molecular dyna- manufacture of solid dispersions, HPMC has been
mics also provides insight into localized interactions shown to inhibit precipitation for a number of different
between molecules, clusters, and the fluid phase and compounds and is arguably the least compound-specific
their relative contributions to the nucleation and precipitation inhibitor (Warren et al., 2010). Nonethe-
crystallization of molecules in complex media (Nishino less, precipitation inhibition by HPMC remains vari-
et al., 2011). Detailed discussion of these concepts is not able. For example, in an examination of the ability
warranted here; however, the importance of cluster of HPMC to inhibit drug precipitation from super-
formation in drug crystallization and the potential to saturated solutions of a series of 24 compounds, pre-
disrupt this process via interaction with excipients, cipitation inhibition was evident for only 12 compounds,
including polymers, is apparent, and a deeper un- whereas for 12 others, little evidence of stabilization
derstanding of the complexities of crystal nucleation was apparent (Vandecruys et al., 2007). Variability in
will likely drive further improvements in formulation the ability of polymers to inhibit precipitation has also
design. been shown with polymers, including PVP and HPMCAS,
At present, however, a priori prediction of the ability where under some circumstances precipitation inhibi-
of an excipient to prevent drug crystallization remains tion may be highly effective (some evidence for the
difficult, and structurally related polymers often show benefits of HPMCAS are shown in Fig. 53), but under
significant differences in precipitation inhibition (Ziller other conditions, effects may be limited (Warren et al.,
and Rupprecht, 1988; Warren et al., 2010). As a result, 2010).
the identification of polymeric carriers that increase Difficulties have also emerged in translating the
dissolution rate and also stabilize supersaturation re- data obtained from screening methods using polymer
mains largely empirical and driven by high-throughput solutions to successful precipitation inhibition during
screening methods. the dissolution of a manufactured dosage form. For
c. Screening for Potential Polymer Precipitation example, vitamin E TPGS 1000 (D-a-tocopheryl poly-
Inhibitors. Several recent papers have used solution- ethylene glycol succinate) and PVPVA-64 effectively
based screening methods to assess the potential for inhibit the precipitation of itraconazole from supersat-
polymeric materials to limit drug precipitation from urated solutions in vitro using solvent-shift methods
supersaturated solutions (Vandecruys et al., 2007; (Janssens et al., 2008). In contrast, dissolution tests
Curatolo et al., 2009; Janssens et al., 2010; Warren conducted using solid dispersion formulations contain-
et al., 2010). The methods used have also been re- ing the same components resulted in drug recrystalli-
viewed by Brouwers et al. (2009) and Warren et al. zation from supersaturated solutions after ;1 h. The
(2010). Although several approaches that assess poly- authors attributed the lack of inhibition of precipita-
meric precipitation inhibitors have been described, tion during the dissolution tests to the presence of seed
the most common approach is the solvent-shift (or crystals in the solid dispersion matrix. Contradictory
cosolvent/solvent quench) method. In solvent-shift results have also been reported for Eudragit 4155F and
methods, supersaturation is stimulated via the intro- PVP (Abu-Diak et al., 2011). In the latter case, PVP
duction of a small volume of a concentrated solution of maintained the stability of supersaturated solutions of
drug in a water-miscible solvent [such as DMSO, celecoxib formed via solvent shift more effectively than
DMA, or dimethylformamide (DFA)] into a receptor Eudragit 4155F, whereas the reverse was true on
medium containing dissolved polymer. Samples are examination of drug dissolution patterns from solid
426 Williams et al.

dispersions containing the same polymers. The im-


proved performance of Eudragit 4155F in the dissolu-
tion experiments was ascribed to superior solubilization
and wetting properties (Abu-Diak et al., 2011).
The variability observed in the performance of
polymeric precipitation inhibitors (PPI) currently pre-
cludes a priori selection of appropriate inhibitors
without experimental testing and is compounded by
a lack of understanding of the mechanisms by which
PPIs act (Warren et al., 2010). This is further com-
plicated by the realization that the most effective
polymers must enhance dissolution and maintain
supersaturation in solution and simultaneously main-
tain drug stability in the solid state in the formulation.
Screening for polymer carriers on the basis of their
ability to support supersaturation should therefore be
used as only one indicator of an appropriate carrier
(Yamashita et al., 2011).

C. Recent Examples of Bioavailability Enhancement


Using Solid Dispersions
Solid dispersions containing drug in the amorphous
form (i.e., types 2 and 3 and types 5 and 6, Fig. 50)
commonly result in faster rates of drug dissolution
compared with crystalline drug, drug-carrier physical
mixtures, and solid dispersions containing drug in the
crystalline form (i.e., types 1 and 4, Fig. 50). Most
recent investigations of the utility of solid dispersion
formulations have therefore focused on the development
of formulations containing drug in the amorphous form
(Newman et al., 2012). Law et al. (2004) have described
PEG 8000–based solid dispersions containing amor-
phous ritonavir and provide evidence of significantly
faster dissolution rates compared with simple physical
mixtures (Fig. 54A) and marked increases in oral Fig. 54. (A) In vitro dissolution in 0.1 N HCl of physical mixture (PM)
containing crystalline ritonavir–PEG at 10:90 and amorphous ritonavir
bioavailability in beagle dogs compared with crystal- in PEG solid dispersions (SD) at various drug:polymer ratios. Dissolution
line drug (Fig. 54B). Increases in drug load, however, was determined using the USP I method (50 rpm, 37°C). Values are
led to phase-separation of drug within the crystalline expressed as means (n $ 2). (B) plasma concentration–time profiles of
ritonavir after a single oral dose of ritonavir formulations to beagle dogs.
PEG carrier, giving rise to decreased dissolution rates Ritonavir was administered in the crystalline form or as the amorphous
and in vivo exposure (Law et al., 2004). Amorphous form in PEG solid dispersions at various drug:polymer ratios. Values are
expressed as means (n 5 7). Adapted from Law et al. (2004).
solid dispersions have also been explored as a potential
formulation strategy for AMG 517, a poorly soluble
vanilloid receptor-1 antagonist (Kennedy et al., 2008). both formulations, the HPMCAS formulation was
Originally formulated as a drug suspension in 10% selected for further testing on the basis of superior
Pluronic F108 (OraPlus), AMG 517 showed good oral performance (28-fold faster intrinsic dissolution com-
bioavailability during preclinical evaluation, although pared with crystalline drug and 2-fold higher than
it was later revealed that this was due to AMG 517 unformulated amorphous drug). Administration of the
forming a cocrystal with sorbic acid, the preservative in HPMCAS solid dispersion to cynomolgus monkeys
OraPlus (Bak et al., 2008) (see Section V.B.5). Un- resulted in a 1.4-fold increase in Cmax and a 1.6-fold
fortunately, the OraPlus formulation was unable to increase in exposure compared with the OraPlus
achieve sufficient exposure using higher AMG 517 suspension (Kennedy et al., 2008).
doses intended for human clinical trials (Kennedy A large volume of data describing the potential
et al., 2008). Solid dispersions were therefore prepared utility of HPMCAS solid dispersions has been de-
by spray-drying solutions of drug and HPMC E5 (a low scribed by Friesen and coworkers (Friesen et al., 2008).
viscosity grade of HPMC) or HPMCAS, both ap- These studies summarized the use of HPMCAS to
proaches resulting in isolation of amorphous drug. generate amorphous solid dispersions for a number of
Although improvements in dissolution were evident for investigational Pfizer compounds. Preclinical exposure
Strategies for Low Drug Solubility 427

data were obtained for more than 100 compounds using attributable to the dissolution of high-molecular-
HPMCAS solid dispersions of amorphous drug, and weight polymeric carriers and segregation of solid
the solid dispersions gave rise to 2- to 40-fold increases dispersion particles resulting in the generation of
in bioavailability compared with crystalline drug. heterogeneous suspensions (Kawakami, 2009).
Twenty-one compounds were evaluated in clinical The ideal properties of a suspending vehicle are the
trials using the same technology, and in all cases, bio- subject of some debate. An aqueous suspending vehicle
availability was improved, with the worst performer containing 1–2% HPMC or 1–5% poloxamer has been
showing a 2-fold increase in bioavailability compared recommended (Nagapudi and Jona, 2008), however,
with crystalline drug (Friesen et al., 2008). Another others caution against the use of additives with solu-
excipient experiencing notable interest is Soluplus bilizing properties as this may encourage rapid dis-
(BASF Corp.), which is a polyvinyl caprolactam–polyvinyl solution of the drug and carrier, increasing the chance
acetate–polyethylene glycol graft copolymer. Soluplus of drug recrystallization (Zheng et al., 2011). In an
solid solutions of BCS class II drugs fenofibrate, attempt to evaluate the impact of predispersion on
danazol, and itraconazole have been prepared by hot- utility, a solid dispersion of the experimental com-
melt extrusion, with each system demonstrating su- pound LCQ789 was recently administered to rats as an
perior in vivo performance in beagle dogs compared aqueous suspension in water, and the data were com-
with the crystalline form of the drug (Linn et al., 2012). pared with administration of the solid dispersion as
The partial lipophilic character of this polymer ex- a powder-filled capsule in dogs (Zheng et al., 2011). In
plains its capacity to increase drug solubilization in both cases, similar rank-order performance was evi-
aqueous media (Hardung et al., 2010). This solubiliza- dent. In this example, the solid dispersion led to
tion capacity may minimize the risk of drug precipita- superior exposure compared with a simple suspension
tion from supersaturated solutions (by decreasing the or cocrystal formulation but lower exposure than that
amount of supersaturated drug). obtained from a microemulsion formulation. The solid
A powerful example of the potential clinical and dispersion, however, showed lower variability in dose-
commercial utility of amorphous solid dispersions is escalation trials and was therefore considered the
provided by Kaletra (Abbott Laboratories) (see Table 26). most favorable approach for ongoing studies (Zheng
Kaletra is a combination of two HIV protease inhibitors, et al., 2011).
ritonavir (50 mg) and lopinavir (200 mg), in a poly-
vinylpyrrolidone/vinylacetate copolymer (PVPVA) solid D. Role of the Carrier and the Drug-Carrier Ratio in
dispersion prepared by hot-melt extrusion (see Section Dictating Drug Release Kinetics from
X.E.1) (Breitenbach, 2002; Gao, 2008). The solid- Solid Dispersions
dispersion formulation was developed in preference to A molecular dispersion of drug within a solid
a liquid-filled gelatin capsule formulation on the basis dispersion is expected to provide the maximum surface
of a diminished food effect in humans and an improved area available for hydration and, therefore, faster
stability profile since the capsule formulation required dissolution rates compared with formulations where
refrigeration (Klein et al., 2007). Sporanox (Janssen drug is present as a fine suspension of amorphous
Pharmaceuticals) provides a second commercial exam- particles. Indeed, the dissolution rate of molecularly
ple of the utility of amorphous solid dispersions. dispersed drug may be almost instantaneous. How-
Although many studies have examined solid disper- ever, where drug dissolution is rapid, the dissolution
sion formulations preclinically in dogs and clinically in rate of the carrier has the potential to limit drug
humans, it is also possible to evaluate system perfor- release (Dubois and Ford, 1985; Craig, 2002; DiNunzio
mance in smaller rodent models (Newman et al., 2012). et al., 2010b). Under conditions of “carrier-controlled
In this case, administration is complicated by the release,” drug dissolution is dependent on carrier
difficulty of administering encapsulated or tableted hydrophilicity and viscosity.
dose forms in small animals. Hard gelatin minicap- Carrier-limited drug release from solid dispersions is
sules have been used (Hasegawa et al., 1985; Yoo et al., typified by a positive correlation between the rate of
2000; Newa et al., 2007; Van Speybroeck et al., 2010b); release of drug and the rate of release or dissolution of
however, the low volume of the rodent stomach may the polymeric carrier (Corrigan, 1986; Craig, 1990;
limit the rate at which formulation is released from the Esnaashari et al., 2005). This occurs more frequently
dosage form. Solid dispersions may be administered as when drug loading is low since this favors the
a suspension, but care must be taken to administer the formation of molecular dispersions (rather than amor-
dose as soon as possible after dispersion to minimize phous suspensions). However, drug dissolution from
the potential for in situ crystallization or precipitation solid dispersions containing suspended amorphous
(Nagapudi and Jona, 2008). For small animals that are drug particles can also be controlled by the rate of
dosed by gavage, formulations are typically dispersed polymer dissolution if dissolution of the amorphous
in the dosing syringe immediately before administra- particles is sufficiently rapid (Bansal et al., 2007; Khan
tion. Complications include increases in viscosity et al., 2011).
428 Williams et al.

The critical processes that occur during carrier-


controlled drug release have been described in detail
by Craig (2002). In situations where polymer dissolu-
tion is fast, drug is in contact with the polymer only
transiently and is liberated into the bulk medium as
a fine molecular dispersion. Increasing carrier viscosity
leads to slower drug release, and an inverse relation-
ship between drug release rate and carrier viscosity
has been described for solid dispersions based on PEG
(Dubois and Ford, 1985; Craig and Newton, 1992;
Shah et al., 1995), PVP (Bansal et al., 2007), HPMC
(Karavas et al., 2001, 2007; Tajarobi et al., 2011) and
Eudragit (Miller et al., 2008; DiNunzio et al., 2010b;
Abu-Diak et al., 2011) High-molecular-weight PVP and
HPMC provide additional barriers to dissolution via
the formation of viscous “gels” that restrict the rate of Fig. 55. Effect of sulfathiazole-to-PVP ratio on the release profile of
water ingress and egress from the matrix (Leuner and sulfathiazole from tablets made from aqueous coprecipitated mixtures of
drug and polymer. The crystalline drug is included as a reference.
Dressman, 2000; Karavas et al., 2001, 2007). This is to Adapted from Simonelli et al. (1976).
some extent analogous to carrier-dependent release of
drug from hydrophilic matrix tablets, albeit in the
latter case, the rate of drug release is intended to occur Craig, 2002). This phenomenon of “drug-controlled
much more slowly. dissolution” has been used to explain the reduced
In addition to increases in dissolution, polymeric release of sulfathiazole from PVP solid dispersions
carriers require properties that enable the stabiliza- (Simonelli et al., 1976) and, more recently, the release
tion of the high-energy amorphous form. The ideal of felodipine and indomethacin from PVP/HPMC solid
polymer properties for amorphous stabilization (high- dispersions (Alonzo et al., 2011). Under these condi-
molecular-weight, high-viscosity, nonhygroscopic) are tions, the rate of drug release is dependent on the
largely contradictory to those that promote rapid re- physical form of the drug in the drug-rich layer. Where
lease. Polymer selection is therefore a tradeoff between drug is present as an amorphous suspension, the drug
properties, including release rate and amorphous drug concentrations that are likely to be achieved during
stabilization (Janssens and Van den Mooter, 2009). dissolution are expected to be supersaturated relative
Drug-release profiles also change with increasing to the solubility of the crystal form but unlikely
drug-carrier ratio, and several studies have shown a (thermodynamically) to exceed the saturated solubility
reduction in drug dissolution rate from solid dispersion of amorphous drug. In contrast, where drug is molec-
with increasing drug-carrier ratios (Corrigan, 1986; ularly dispersed, rapid initial dissolution may lead to
Veiga et al., 1993; Kapsi and Ayres, 2001; Verreck supersaturation relative to both crystal and amor-
et al., 2003b; Law et al., 2004; Karavas et al., 2005; phous drug solubilities. Consistent with this sugges-
Janssens et al., 2008; Konno et al., 2008; Lakshman
tion, felodipine-HPMC solid dispersions assembled at
et al., 2008; Abu-Diak et al., 2011; Alonzo et al., 2011;
low drug-carrier ratios of 10:90 (that favor drug-
Dontireddy and Crean, 2011; Ghosh et al., 2011; Han
polymer miscibility) show very fast initial rates of dis-
et al., 2011; Khan et al., 2011) with one example re-
solution, and dissolved drug concentrations exceed the
produced in Fig. 55 (Simonelli et al., 1976). The de-
calculated solubility of amorphous drug (Alonzo et al.,
crease in dissolution is usually attributed to a
reduction in the quantity of drug in the molecularly 2011). Similar observations have been made during
dispersed form since increasing drug load reduces the dissolution testing of felodipine-PVP and indomethacin-
intermolecular distance between drug molecules and PVP solid dispersions at the same 10:90 drug-carrier
the quantity of stabilizing polymer, both of which favor ratio. However, dynamic light scattering experiments
phase separation of amorphous/crystalline particulate have subsequently revealed that the concentration of
drug in the carrier (Vasanthavada et al., 2005; Qian felodipine “in solution” after release from the HPMC
et al., 2010a). solid dispersions also include drug in the form of
At high drug-carrier ratios, the reduction in the submicron-sized particles, which absorb in the same
relative quantity of polymer present may also lead to region of the UV spectrum as dissolved drug. The true
a decrease in drug wetting and the potential for concentration of drug in solution is therefore lower and
incongruent release of drug and carrier, with the latter more consistent with the amorphous drug solubility
resulting in the formation of a carrier-depleted, drug- (Alonzo et al., 2011). Nonetheless, solid dispersions
rich solid phase that is at increasing risk of phase comprising drug molecularly dispersed throughout the
separation or crystallization (Higuchi et al., 1965a; carrier are likely to show maximal rates of drug release
Strategies for Low Drug Solubility 429

compared with formulations containing suspended More recently, hot-melt extrusion (HME) has grown
amorphous drug and are typically preferred. in popularity as it appears to address many of the
limitations of simple fusion methods (Breitenbach,
E. Common Methods to Manufacture 2002; Crowley et al., 2007; Repka et al., 2007). In
Solid Dispersions HME, a mixture of drug and carrier(s) is introduced via
a hopper into an extruder that contains a rotating
1. Melting/Fusion. One of the simplest methods of
screw or auger inside a heated barrel. The extruder is
solid dispersion formation is via heating mixtures of
rapidly heated to lower the viscosity of the drug-carrier
drug and carrier above the melting or glass transition
blend. The rotating screw, in combination with heat-
temperature, followed by solidification by rapid cool-
ing, results in effective deaggregation of drug particles
ing. Adequate mixing to form a homogeneous molecu-
and incorporation into the molten carrier. The mixture
lar solution is critical to the physical stability and
is subsequently forced through a die to produce an
performance of solid dispersions formed via fusion
extrudate of uniform shape (Crowley et al., 2007). The
techniques since this avoids formation of the drug-rich
residence time of the blend in the extruder is short
or polymer-rich microenvironments that promote (;2–5 min), limiting thermal instability (Breitenbach,
phase separation and reduce dissolution rate. Rapid 2002). HME can be used in batch mode or adapted for
cooling is also essential as conditions where the cooling continuous production (Breitenbach, 2002).
rate exceeds the rate of crystallization are required to Proprietary HME methods include the Meltrex
facilitate isolation of drug in the more soluble amor- technique developed by SOLIQS (Abbott GmbH & Co.
phous form (see Section X.F.1). Where phase separa- KG), which is used in the production of Kaletra and
tion does occur, rapid cooling reduces the time Isoptin-SRE tablets (both developed by Abbott Labo-
available for particle growth and reduces the particle ratories), and the KinetiSol dispersing (KSD) technol-
size of the (amorphous or crystalline) suspended ogy, which was recently developed at the University of
material. Texas and now owned by DisperSol Technologies LLC
The first reported solid dispersions were formed (Austin, TX). KSD uses a series of rapidly rotating
by fusing sulfathiazole and urea above the eutectic blades that generate a level of kinetic and thermal
temperature, prior to rapid cooling in an ice bath energy that is sufficient to promote rapid drug-carrier
(Sekiguchi and Obi, 1961). Faster rates of cooling were fusion without the need for an external heat source
later achieved by pouring the melt onto stainless steel (DiNunzio et al., 2010a,b). The shorter time required to
plates to accelerate heat loss (Sekiguchi et al., 1964; achieve an intimate mixture of drug with carrier
Chiou and Riegelman., 1970). Ice bath and liquid compared with conventional HME may be of particular
nitrogen quenching remain widely used cooling tech- significance for drugs that are thermally labile. Indeed,
niques (Dordunoo et al., 1991; Saers et al., 1993; HPMC-itraconazole solid dispersions have been pre-
Hirasawa et al., 1999; Law et al., 2001). Melts may also pared using KSD in ,15 s, whereas conventional HME
be spray-cooled (spray-congealing). In this process, requires .5 min (DiNunzio et al., 2010b). In the same
melts are sprayed into a cooled chamber that is study, the KSD solid dispersions were also shown to be
continuously perfused with chilled air, allowing the more stable than those formed by conventional HMEs,
formation of spherical drug/carrier particles that can and the increase in physical stability was attributed to
be directly filled into a capsule or compressed into more homogeneous mixing and, therefore, a lower risk
a solid tablet (Sancin et al., 1999; Fini et al., 2002). of phase separation. The ability to heat rapidly and melt
Drug release rates from solid dispersions typically drug-carrier blends using KSD has the additional
increase with increasing cooling rate since this in- advantage of reducing the need for plasticizers. Plasti-
creases the likelihood of drug isolation in the noncrys- cizers are commonly added to facilitate melting during
talline form and minimizes the size of any suspended HME, but they may be detrimental to the long-term
particles (McGinity et al., 1984; Save and Venkitacha- stability of solid dispersions beause of their propensity
lam, 1992). In contrast, reports of slower dissolution to increase molecular mobility (Lakshman et al., 2008;
rates with an increased cooling rate are also apparent DiNunzio et al., 2010a). Finally, HME and KSD also
(Saers et al., 1993; Doshi et al., 1997), illustrating the allow the formation of solid dispersion for high melting
unpredictable nature of crystallization (Craig, 1990; polymers such as HPMC and HPMCAS (Verreck et al.,
Saers et al., 1993). 2003b; Six et al., 2005; DiNunzio et al., 2010b; Hughey
Limitations to hot-melt methods for preparing solid et al., 2010), which are most commonly manufactured
dispersions include the possibility of poor drug-carrier using spray-drying or other solvent-removal methods
miscibility at elevated temperatures (Forster et al., owing to the high temperatures required to melt the
2001a) and the potential for thermal degradation of polymer (Tg . 120°C) (Gilis et al., 1995; Jung et al.,
drug where it is necessary to hold the drug-carrier 1999; Han et al., 2011; Won et al., 2005).
mixture at elevated temperatures for prolonged peri- 2. Solvent Evaporation. Before the development of
ods (Goldberg et al., 1965). hot-melt extrusion, the thermal degradation concerns
430 Williams et al.

surrounding traditional melting methods dictated the a dichloromethane-ethanol solution (Janssens and Van
use of solvent evaporation techniques in the manufac- den Mooter, 2009). Etravirine-HPMC solid dispersions
ture of solid dispersions, particularly when using high (Intelence; Tibotec) are also prepared by spray-drying
Tg carriers (Leuner and Dressman, 2000). The temper- (Scholler-Gyure et al., 2009). More efficient solvent
atures required to evaporate commonly used organic removal enhances product performance since it limits
solvents are usually in the order of ;25–65°C (Mas- the time available for drug and polymer to migrate
ters, 1991) and lower than those required during melt/ and phase separate. Thermal analysis of itraconazole-
fusion methods. Solvent evaporation was therefore Eudragit E100 solid dispersions has shown that in-
preferred for thermally labile drugs or for carriers with creasing the speed of solvent removal by spray-drying
high Tg values, for example HPMC and PVP (Padden results in improved drug-polymer miscibility compared
et al., 2011). with equivalent cast-film solid dispersions that rely on
During the manufacture of solid dispersions via solvent evaporation (Janssens et al., 2010). In this
solvent evaporation, drug and carrier are dissolved in example, the use of spray-drying allowed higher drug
a common solvent before mixing and solvent evapora- loading without increasing the risk of phase separation.
tion under elevated temperatures or reduced pressure. In summary, solvent evaporation methods allow the
This is typically achieved using spray-drying (Cal and use of lower temperatures than traditional melt/fusion
Sollohub, 2010; Sollohub and Cal, 2010). Solvents techniques but are complicated by the identification of
include chloroform, dichloromethane, methanol, etha- solvent systems with appropriate solvent capacity for
nol, acetone, and methylene chloride. Solvents may be both carrier (e.g., polymer) and drug. Where drug/
used in combination and may also include small carrier solubility is low and large batches are required,
amounts of water to enhance drug or polymer solubility solvent volumes are high, resulting in significant en-
in the solvent mixture (Sinha et al., 2009; Padden vironmental, cost, and procurement issues (Serajuddin,
et al., 2011). 1999). Solvent removal is also critical, and since the rate
The solvent must be completely removed after of solvent removal influences molecular miscibility and
formation of the solid dispersion since residual solvents structure within the carrier, solid dispersion perfor-
carry a toxicity liability and also plasticize the drug- mance can vary significantly with small changes in manu-
polymer mixture. Plasticization enhances mobility and facturing conditions.
increases the likelihood of phase separation and 3. Solvent Evaporation Using Supercritical Fluids.
crystallization. Solvent removal is usually achieved Supercritical fluids (SCFs) can be used to enhance the
under vacuum at elevated temperature and may also efficiency of solvent evaporation during the manufac-
require desiccation. For example, methanol removal ture of solid dispersions. In a manner analogous to the
from carbamazepine-PVP K30 solid dispersions has advantages of spray-drying and freeze-drying, SCF
been reported to occur under vacuum in a rotary evap- methods facilitate rapid removal of solvent (Pasquali
orator at 40°C over a 24-h period (Sethia and Squillante, et al., 2006). SCF techniques also use lower volumes of
2004), ethanol from indomethacin-HPMC/ethylcellulose organic solvents, minimize disposal costs, and reduce
mixtures under vacuum at 35°C by rotary evaporation the potential for solvent carryover into the final pro-
and then further drying for 12 h in a desiccator (Ohara duct. SCFs that are commonly used in pharmaceutical
et al., 2005), and chloroform from ofloxacin-PEG 4000/ processing and the different solvent and antisolvent
20000 dispersions under reduced pressure using a vac- methods that may be used with SCFs to isolate
uum dryer at 40°C before further drying in a vacuum pharmaceutical products are discussed with reference
oven (at room temperature) for 24–48 h (Okonogi et al., to nanoparticle preparations in Section IX. The prin-
1997). More efficient drying is possible at elevated ciples of the use of SCFs to promote solid dispersion
temperatures (Chiou and Riegelman., 1970), but in- manufacture are similar to those used during nano-
creases in temperature also increase molecular mobility particle or microparticle isolation, the major difference
and, therefore, the likelihood of amorphous to crystal- being the addition of larger quantities of polymer, such
line transitions. Some solvents are easier to remove that the polymer acts as a matrix rather than a
than others (Bitz and Doelker, 1996), however, solvent stabilizer. In both cases, SCFs are now most commonly
choice is largely driven by toxicity burden and the used as an antisolvent (SAS techniques) rather than
capacity to dissolve both drug and polymer rather than the primary solvent because of the low solubility of
by the vaporization pressure. many drugs in SCFs (Jun et al., 2005; Won et al., 2005;
More effective approaches to the removal of solvent Majerik et al., 2007; Li et al., 2011b).
include freeze-drying (Otsuka et al., 1993; Van Drooge Supercritical solvent methods have been used to
et al., 2004), spray-drying (Jung et al., 1999; Van den form solid dispersions for drugs, including carbamaze-
Mooter et al., 2001; Weuts et al., 2005), and fluidized pine (Moneghini et al., 2001), felodipine (Won et al.,
bed granulation (Ho et al., 1996). The Sporanox solid 2005), cefuroxime axetil (Jun et al., 2005), phenytoin
dispersion (itraconazole-HPMC; Janssen Pharma- (Muhrer et al., 2006), piroxicam (Wu et al., 2009), and
ceuticals), for example, is prepared by spray-drying from ketoprofen (Manna et al., 2007). SCF-enabled solid
Strategies for Low Drug Solubility 431

dispersions may provide advantages in both dissolution F. Physical Stability of Amorphous Solid Dispersions
rate and processability. For example, carbamazepine- Physical stability is a critical issue during product
PEG 8000 solid dispersions prepared by gas anti- development for solid dispersion formulations. For
solvent (GAS) precipitation result in improved flow amorphous solid dispersions in particular, complica-
properties and lower cohesion compared with those tions arise from the potential for amorphous drug to
formed by conventional solvent evaporation (Sethia revert back to the more stable (but less soluble)
and Squillante, 2002). Piroxicam-PVP solid dispersions crystalline form. Where drug is molecularly dispersed
prepared by precipitation with compressed antisolvent in the carrier [i.e., type 3 (crystalline carrier) or type 6
(PCA) methods also result in small increases in
(amorphous carrier) systems; Table 26] transformation
dissolution rate compared with the equivalent spray-
to an amorphous suspension (i.e., type 2/5) or crystal-
dried solid dispersion. In the latter example, piroxicam
line suspension (i.e., type 1/4)] is likely and must be
was present as an amorphous dispersion in both PCA
controlled to provide appropriate product shelf life.
and spray-dried solid dispersions, and the faster dis-
As the amorphous to crystalline drug conversion is
solution rate from the PCA product was ascribed to
usually associated with a decrease in drug dissolution
the presence of smaller suspended particles in the
and a decrease in oral bioavailability, efforts to better
PCA-derived matrix (Wu et al., 2009). In contrast, solid
understand and minimize physical transformations are
dispersions prepared by antisolvent SCF techniques
increasing. It is now widely recognized that an un-
have in some cases been shown to exhibit slower dis-
derstanding of the thermodynamic drivers (i.e., the free
solution rates compared with the equivalent spray- or
energy difference between the amorphous and the
freeze-dried product as a result of increased drug
more ordered crystalline phase) and the kinetic drivers
crystallinity in the manufactured product (Corrigan
(i.e., the level of mobility or freedom to move within the
and Crean, 2002; Majerik et al., 2007; Badens et al.,
solid state) of crystallization is central to the design of
2009).
4. Electrospinning and Microwave Irradiation. solid dispersions that show adequate physical stability
Electrospinning, or electrostatic spinning, can be used over pharmaceutically relevant periods. A complete
to generate very long nanofibers of uniform diameter understanding of the properties of pharmaceutical
and has the potential to assist in the manufacture of glasses, however, remains elusive. Indeed, in 2005,
any polymer-based pharmaceutical product (Ignatious on the 125th anniversary of the journal Science,
et al., 2010). In the manufacture of solid dispersions by a section was devoted to 125 questions that point to
electrospinning, a charged drug-polymer melt or gaps in our basic knowledge, and among these was the
solution is passed through a nozzle and subjected to question, What is the nature of the glassy state?
a high voltage (5–30 kV direct current). At a critical (Anonymous, 2005). The nature and properties of the
point, the electrostatic repulsive forces that result from glass state continue to be a topic of extensive research
the applied charge overcome the surface tension in the in physics and physical chemistry (Prevosto et al.,
fluid, and a fine stream of fluid (typically with a 2008; Wondraczek and Mauro, 2009; Ngai et al., 2010;
diameter that is orders of magnitude smaller than the Angell, 2011; Berthier and Biroli, 2011; Johari, 2011a,
nozzle diameter) is ejected from the fluid surface. The b; Capaccioli et al., 2012), and it remains a challenge
nanofibers that are produced have very small diame- for pharmaceutical scientists to keep abreast of the
ters (;100 nm) and very high surface area-to-volume advances made.
ratios (Ignatious and Baldoni, 2001; Ignatious and Detailed descriptions of transitions in the glass state
Sun, 2004), resulting in rapid solidification via cooling and the mechanisms by which amorphous materials
or solvent evaporation. The efficiency of solvent evap- revert to the crystalline state are beyond the scope
oration also minimizes the potential for solvent re- of this review, and the interested reader is directed
sidue. The dimensions of electrospun solid dispersions toward Byrn et al. (1999), Hancock and Zografi (1997),
can be controlled via the magnitude of the applied Bhattacharya and Suryanarayanan (2009), and Johari
voltage to provide customized rates of drug release and Shanker (2010) for detailed information on molec-
(Fridrikh et al., 2003; Verreck et al., 2003a). ular mobility, relaxation, and crystallization. Nonethe-
Finally, microwave irradiation has recently been less, a relatively brief discussion of the formation
used to promote solid dispersion formation. Micro- of pharmaceutical glasses and the mechanisms of
waves rapidly achieve uniform heating in polymeric (in)stability of amorphous material is required and
materials, and short periods of (,5 min) microwave is provided in the following as critical background
irradiation have been used to manufacture solid dis- to the design of (physically) stable pharmaceutical
persions of tibolone-PEG (Papadimitriou et al., 2008), preparations.
nimesulide-Gelucire 50/13 and nimesulide-Poloxamer 1. Structural Changes at the Glass Transition Temp-
188 (Moneghini et al., 2009), atorvastatin calcium-PEG erature. Amorphous solids are produced from the crys-
6000 (Maurya et al., 2010), and itraconazole-TPGS talline form using high-energy processing techniques
(Moneghini et al., 2010). such as melt extrusion, melt quenching, milling, grinding,
432 Williams et al.

or compaction or by direct isolation from a supersaturated system that enters a supercooled liquid phase must at
solution via precipitation, freeze-drying, or spray-drying some point either enter into the glassy state or crys-
(Hancock and Zografi, 1997). All amorphous solids lack tallize since these phase changes are necessary to
the long-range order evident in crystalline solids, and avoid an “entropy crisis,” in which the extrapolated
in pharmaceutical applications, solid dispersions con- entropy of the supercooled liquid (i.e., the dashed green
taining amorphous drug typically comprise drug in line) falls below that of the crystalline solid. This ther-
the “glass” state. Amorphous glasses are amorphous modynamic crisis (commonly referred to as Kauzmann’s
solids that convert to the liquid state on heating paradox) is avoided through glass formation. The Kauzmann
through a glass-liquid transition temperature and are temperature (TK) may be interpreted as the lowest
formed during rapid cooling of the molten state. This possible Tg for glass formation.
is summarized in Fig. 56, which describes changes in In the glassy state, entropy, enthalpy, and volume
molecular volume with changes in temperature. An are higher than in the crystalline state, and amorphous
understanding of the events that occur during the solids are therefore thermodynamically unstable and
liquid-glass transition (on cooling) also provides a liable to relax and recrystallize over extended periods.
background to the issues surrounding the physical (in) Over shorter times, molecular relaxation (and there-
stability of solid dispersions containing amorphous fore crystallization) is inhibited by the dramatic reduc-
tion in molecular mobility below the Tg, resulting from
drug and the ability of excipients such as polymers to
the increase in the viscosity of the glass (Davies and
enhance physical stability.
Jones, 1953; Angell, 1995a). The increase in viscosity
The solid blue line in Fig. 56 traces the marked
in turn limits molecular diffusion and reorientation
reduction in system volume that occurs as a melt is
into the position required to form a crystal lattice.
cooled below the melting temperature (Tm), leading to
2. Glass-Forming Potential. All materials are the-
crystallization of the thermodynamically stable solid. In
oretically capable of forming a glass in the manner
contrast, the solid green line traces the change in system
described in the preceding section (Turnbull, 1969). In
volume under rapid cooling. Under these conditions, the
practice, however, many relax rapidly on cooling and
rapid rate of cooling precludes molecular relaxation crystallize rather than forming kinetically stable
and crystallization below the melting temperature, and glasses (Turnbull, 1969; Yu, 2001). Drugs showing
instead, the material enters a supercooled liquid phase. these characteristics are said to have low glass-forming
Further cooling of the metastable supercooled liquid potential and form glasses only under rapid rates of
leads to a critical point, the glass transition tempera- cooling (Kaushal and Bansal, 2008; Baird et al., 2010).
ture (Tg), where, if the rate of cooling is faster than the The minimum rate of cooling at which a material
rate of molecular relaxation, the material enters vitrifies to form a glass is the critical cooling rate (Rcrit),
a glassy, amorphous state. The thermodynamic basis and this value may be used to compare differences in
for the glass transition was proposed by Kauzmann glass-forming ability across a drug series (Baird et al.,
(1948). In brief, Kauzmann (1948) suggests that a 2010). Compounds with high Rcrit values are poor
glass-formers, whereas low Rcrit values imply relatively
facile glass formation.
An alternative descriptor for glass-forming potential
and crystallization tendency is the fragility of a mate-
rial (Angell, 1995a,b). Fragility provides an indication
of the rate of structural relaxation and is evidenced by
the viscosity changes on cooling toward Tg. Fragile
materials show significant deviations from Arrhenius
behavior on cooling, leading to rapid rates of relaxation
and a collapse of liquid structure with decrease in
temperature (Angell, 1995a). This propensity for rapid
structural relaxation limits stabilization of an amor-
phous glass. In contrast, the viscosity of “strong”
liquids declines in closer accordance with Arrhenius
relaxation behavior and promotes glass formation
(discussed in more detail in Section X.G.3).
Fig. 56. Schematic depicting the change in molecular volume as The rapid structural relaxation of weak glass-
a function of decreasing temperature. Equivalent changes in entropy formers on cooling is also manifest in high heat–
and enthalpy on cooling are also evident. Illustrative melting tempera-
ture (Tm), glass transition temperature (Tg), and Kauzmann temperature capacity changes (DCp) on passing through the Tg
(TK) of the material are shown. Rapid cooling of a melt results in the (Angell and Tucker, 1980; Kerc and Srcic, 1995;
generation of a supercooled liquid. At the glass transition temperature
(Tg), the rate of cooling exceeds the rate of relaxation, and the material Hancock and Zografi, 1997; Shamblin et al., 1999).
forms a glass. The significant molecular reorganization that occurs in
Strategies for Low Drug Solubility 433

weak glass-formers as temperature changes dictates increasing the Tg of the system and reducing molecular
large heat capacity changes (i.e., high DCp) to dissipate mobility (Ford and Rubinstein, 1979; Yoshioka et al.,
energy. In contrast, limited molecular reorganization 1995; Matsumoto and Zografi, 1999; Van den Mooter
on cooling is evident in strong glass-formers, despite et al., 2001; Six et al., 2004; Karavas et al., 2005;
changing temperature, and as a result, systems exhibit Konno and Taylor, 2008). The use of high-molecular-
much smaller DCp values. DCp therefore provides an weight polymers as vehicles for solid dispersion for-
additional indicator of glass-forming potential, with mation, therefore, provides an opportunity to both
high DCp values suggesting weak glass-forming increase the likelihood of glass formation and promote
potential. the stability of the amorphous system formed through
3. Polymer Effects on the Glass Transition Temp- increased Tg (the capacity for polymers to stabilize
erature. The high molecular viscosity of macromolec- amorphous drug via other mechanisms is further
ular polymers, coupled with a marked reduction in discussed in Section X.G).
system viscosity and molecular mobility on cooling, Zografi and coworkers proposed that mobility within
results in ready stabilization of the glass phase. The an amorphous system was reduced sufficiently to allow
high viscosity of many polymers also results in high Tg acceptable physical stability assuming storage at
values, in excess of that of most drugs (Matsumoto and temperatures 50°C below the Tg (Hancock et al.,
Zografi, 1999; Lakshman et al., 2008; Padden et al., 1995; Hancock and Zografi, 1997). The Tg250°C
2011). For example, the Tg of PVP K30 is 168°C, and suggestion stemmed from a series of papers that
the Tg of HPMC is ;170°C. The high Tg of these examined the recrystallization of indomethacin alone
polymers dictates that homogeneous dispersions of and after the formation of a number of PVP solid
drug and polymer have Tg values that are elevated dispersions. In the absence of polymer, indomethacin
relative to the Tg of drug alone (assuming Tgdrug , recrystallization occurred in less than 6 weeks at
Tgpolymer) and, therefore, that polymeric formulations storage temperatures 20°C below the Tg. When the
have the potential to stabilize molecular transitions in drug was molecularly dispersed in PVP solid disper-
the amorphous state. The ability of a high Tg polymer sions, the difference between the Tg of indomethacin
to raise the apparent Tg of a drug in a solid dispersion is in the formulations and the storage temperature
shown graphically in Fig. 57, where gradual increases increased to 40–50°C. Under these conditions, recrys-
in PVP content in indomethacin-PVP binary mixtures tallization of indomethacin was suppressed (Yoshioka
leads to a progressive increase in the Tg of the system et al., 1994, 1995).
(Yoshioka et al., 1995). For drugs with low glass- Knowledge of the Tg of a drug-polymer solid dis-
forming potential, fusion with a polymer provides persion therefore provides insight into the likely
a means of kinetic stabilization (and, therefore, iso- physical stability of an amorphous solid dispersion,
lation) of the noncrystalline amorphous glass that might and solid dispersions of high Tg are expected to show
not be possible in the absence of polymer because of greater physical stability. Tg may be determined
rapid molecular relaxation and crystallization. experimentally using DSC or other calorimetric meth-
Even where it is possible to isolate amorphous drug ods (Royall et al., 1998; Weuts et al., 2003; Baird and
in the absence of excipient (i.e., for drugs with higher Taylor, 2012). Alternatively, the Gordon-Taylor equa-
glass-forming potential), polymer incorporation usually tion (eq. 52) can be used to predict the Tg of an ideally
increases the physical stability of amorphous drug by mixed solid solution from the sum of the weight
fraction (W) and Tg values of the individual compo-
nents of the mixture (Gordon and Taylor, 1952):
 
w1 Tg1 1 Kw2 Tg2
Tgmix 5 ð52Þ
ðw1 1 Kw2 Þ

The parameter K in the Gordon-Taylor equation is a


constant and may be approximated from the densities
(r) of the components in the amorphous state according
to the Simha-Boyer rule (Simha and Boyer, 1962):

r  1 Tg1
K5 ð53Þ
r  2 Tg2

In eq. 52 and eq. 53, subscripts 1 and 2 are used to


Fig. 57. Plot of glass transition temperature (Tg) versus weight percent represent individual components. Analogous equations
of PVP for indomethacin-PVP coprecipitates. The symbols represent data
points, and the solid line represents the fit to the Gordon-Taylor equation may be derived for more complex mixtures (i.e., those
(eq. 52). Adapted from Yoshioka et al. (1995). containing more than two components) (Van den
434 Williams et al.

Mooter et al., 2001); alternatively, an approximation of the potential to hydrogen bond with hydrophilic
total properties can be obtained by using the compo- polymers, thereby disrupting the drug-polymer inter-
nents with the highest and lowest Tg values. actions that may be critical to the physical stability of
A comparison between measured and predicted Tg a solid dispersion (Konno and Taylor, 2008; Rumondor
(using the Gordon-Taylor equation) is shown in Fig. 57, et al., 2009b; Rumondor and Taylor, 2010). Water
and good agreement has also been reported elsewhere absorption on storage can be minimized by appropriate
(Forster et al., 2001a; Konno and Taylor, 2006; van storage conditions, packaging, or polymer choice where
Drooge et al., 2006; Janssens and Van den Mooter, less hydrophilic, less hygroscopic polymers typically
2009; Weuts et al., 2011). Deviations between experi- reduce water uptake.
mental Tg values and those predicted from the Gordon-
Taylor equation usually indicate nonideal mixing in G. Factors Affecting Drug Crystallization from
the solid dispersion (Six et al., 2003a) resulting from Solid Dispersions
poor drug-polymer miscibility or plasticization by The use of high Tg polymers to generate high Tg solid
residual water or other solvents (van Drooge et al., dispersions is a common and often successful mecha-
2006; Patterson et al., 2008). Poor mixing also results nism of stability enhancement. The Tg of a solid dis-
in an increased propensity for phase separation on persion, however, is not always predictive of physical
storage (see X.G.5). In contrast, predicted Tg values stability. Thus, polymer addition may decrease drug
that are consistent with experimental values are crystallization within a solid dispersion but has little
usually indicative of good mixing, although this is not or no effect on Tg (Yoshioka et al., 1995; Miyazaki et al.,
always the case (Marsac et al., 2009; Baird and Taylor, 2007; Konno et al., 2008) or, in some cases, may reduce
2012). Positive deviations from Gordon-Taylor behav- crystallization and simultaneously decrease Tg (Khougaz
ior (where experimental Tg values are greater than and Clas, 2000; Law et al., 2001; Janssens et al., 2010).
estimated values) usually indicate the presence of A broader examination of the physical instability of
molecular interactions between drug and polymer amorphous drug in solid dispersions, beyond that dic-
(Shen and Torkelson, 1992; Khougaz and Clas, 2000). tated by Tg, is therefore warranted.
4. Moisture Effects on Glass Transition Temperature. Most studies that have evaluated the stabilllity of
Although solid dispersions with increased Tg provide solid dispersions have focused on the properties of the
for enhanced stability of incorporated amorphous drug, formulation as a driver of drug stability. These include
under some circumstances, solid dispersions with high examinations of the effects of intermolecular interac-
Tg values might still show evidence of crystallization. A tions between drug and excipient, excipient-mediated
common explanation for this phenomenon is the inhibition of crystal nucleation, changes to molecular
plasticizing effect of moisture adsorption during stor- mobility, and differences in mixing and miscibility.
age (Zhou et al., 2002; Miyazaki et al., 2004; Rumondor These are addressed later herein. In addition, the
and Taylor, 2010). Water has a low Tg (2137°C) and is intrinsic properties of the drug (rather than the for-
a very effective plasticizer (Roos, 1997). As a rule of mulation) also play a role in determining the potential
thumb, 1% adsorbed water increases molecular mobil- for drug crystallization. Several authors have sug-
ity by two orders of magnitude and reduces Tg by ;10° gested that the propensity for drug crystallization from
C (Andronis and Zografi, 1998). The presence of water, the unformulated amorphous form may provide a use-
therefore, facilitates drug relaxation to the more stable ful measure of the potential for crystallization from
crystalline state. Vasanthavada et al. (2005) described a formulated solid dispersion (Marsac et al., 2006a;
a series of griseofulvin-PVP solid dispersions with very Friesen et al., 2008; Janssens and Van den Mooter,
high Tg values (.125°C) but that showed evidence of 2009; Weuts et al., 2011). This is usually assessed via
phase separation and drug recrystallization after differences in the ease of glass formation (i.e., glass
storage at high relative humidity (40°C/69% relative forming ability) and differences in the stability of the
humidity). Moisture-induced instability has also been pure drug glass to crystallization (Aso et al., 2000; Six
reported for PVP solid dispersions containing felodipine et al., 2001; Marsac et al., 2006a; Kaushal and Bansal,
and loperamide (Weuts et al., 2005; Konno and Taylor, 2008; Baird et al., 2010; Van Eerdenbrugh et al.,
2006, 2008; Marsac et al., 2008). In this case, sensi- 2010a; Weuts et al., 2011).
tivity to moisture was ascribed to the hygroscopic As described in Section X.F.2, drugs with low glass-
nature of PVP as similar solid dispersions containing forming potential relax rapidly on cooling, and this is
the less hygroscopic polymers HPMC and HPMCAS manifest in high DCp at Tg. The predisposition of
showed reduced instability on exposure to moisture fragile liquids to structural relaxation facilitates
(Konno and Taylor, 2008). Tablet excipients, including crystallization, decreases the likelihood of glass forma-
fillers, may also be hygroscopic, and incorporation of tion, and reduces glass stability (Angell, 1995b; Mao
such excipients into solid dispersion formulations et al., 2006b; Kaushal and Bansal, 2008; Baird et al.,
should be approached with caution (Bruce et al., 2010). Where glass formation is encouraged by the
2010). Absorbed water has the added complexity of addition of a high Tg polymer, the intrinsic molecular
Strategies for Low Drug Solubility 435

properties of drugs with low glass-forming potential molecular mobility gave a better indication of crystal-
may therefore reduce the physical stability of the lization tendency. Others have also suggested that
amorphous material in formulated solid dispersions, in increases in configurational entropy may be used to
some instances despite favorable changes to other provide an indication of increased stability of amor-
common indicators of stability, such as Tg (Weuts et al., phous materials (Zhou et al., 2008; Graeser et al., 2009;
2005; Miyazaki et al., 2007; Friesen et al., 2008). Baird and Taylor, 2012). In contrast, Johari and
The tendency for drugs to crystallize out of high- Shanker (2010) question the use of thermodynamic
energy amorphous formulations is predicated by the descriptors (such as heat capacity and melting en-
excess volume, enthalpy, entropy, and free energy thalpy) to predict the kinetic stability of amorphous
carried by amorphous glasses compared with the stable solids. This is described in more detail in Section X.G.3.
crystal state. To initiate crystallization, however, an 1. Drug-Polymer Intermolecular Association. Drug-
activation energy barrier must be overcome that polymer interactions, such as hydrogen bonding, de-
reflects the kinetic limitations to crystallization. This crease the ease of recrystallization by reducing drug
is the energy required for molecules to diffuse and mobility and by limiting the drug-drug molecular in-
appropriately align such that nucleation and the teractions that precede nucleation and crystallization
formation of a new solid phase are possible (James, (Khougaz and Clas, 2000; Weuts et al., 2003; Miyazaki
1985). The stability of amorphous drug is a function of et al., 2006; Bhugra and Pikal, 2008). Hydrogen bonds
both the inherent thermodynamic liability of the between indomethacin and PVP, for example, limit
amorphous form and the kinetic barriers to crystalliza- indomethacin crystallization by preventing the for-
tion provided by the viscosity of the glass and mation of drug-drug dimers through adjacent carbox-
supplemented by the addition of excipients, such as poly- ylic acid groups—an interaction likely to be the initial
mers (Van den Mooter et al., 1998; Konno and Taylor, step in indomethacin crystallization (Taylor and
2006; Bhugra and Pikal, 2008; Marsac et al., 2008). Zografi, 1997). Suppression of indomethacin crystalli-
Differences in the thermodynamic properties of drug zation in the presence of PVP has also been reported
in the amorphous versus the crystalline form are at polymer levels (5%) that are too low to have a
typically described in terms of configurational or significant impact on Tg, suggesting other mecha-
conformational properties (i.e., differences in enthalpy, nisms of stabilization (Yoshioka et al., 1995; Matsu-
entropy, and free energy resulting from differences in moto and Zografi, 1999). PVP hydrogen bonds with
structure) and are commonly calculated from the many drug compounds in solid dispersions, and these
configurational heat capacity (i.e., the difference in Cp provide the opportunity for increases in physical
between the amorphous and crystalline states). stability (Van den Mooter et al., 1998; Matsumoto and
A detailed description of the use of configurational Zografi, 1999; Tantishaiyakul et al., 1999; Damian
thermodynamics to describe differences in amorphous et al., 2002; Weuts et al., 2003; Jun et al., 2005; Shinde
drug stability is beyond the scope of the current et al., 2008; Lima et al., 2011). PVP has two functional
discussion; however, two recent studies exemplify the groups (5N- and C5O) with the capacity to accept
potential role of thermodynamic liabilities and kinetic hydrogen bonds, although steric hindrance around the
barriers in dictating appropriate stability for amor- nitrogen generally favors intermolecular hydrogen
phous drug. Marsac et al. (2006a) examined the bonding via the carbonyl group (Van den Mooter et al.,
thermodynamic and kinetic drivers of drug stability 1998). Drug-polymer hydrogen bonding has also been
to crystallization from felodipine and nifedipine amor- shown for several other drug-polymer combinations, for
phous glasses. Nifedipine crystallized more readily example, carbamazepine-PEG 4000/6000 (Doshi et al.,
than felodipine, and this was attributed to a larger 1997), benzidazole-PEG 6000 (Lima et al., 2011),
free-energy difference between the amorphous and the cefuroxime-HPMC (Jun et al., 2005), and NVS981-
crystalline forms, resulting in a greater thermodynamic HPMC phthalate (Ghosh et al., 2011). The absence of
driver for crystallization. Differences in free energy a drug-polymer intermolecular interaction within
were largely derived from differences in enthalpy. In a solid dispersion, however, does not necessarily
contrast, no differences were evident in many deter- preclude the ability of a polymer to stabilize the
minants of the kinetic barriers to crystallization, in- amorphous form (Van den Mooter et al., 2001; Caron
cluding Tg and mobility (although the activation energy et al., 2010). The extent to which drug and polymer
for nifedipine crystallization was lower than that of interact also determines the level of mixing and,
felodipine). Zhou et al. (2002) similarly evaluated the therefore, miscibility. The importance of miscibility to
crystallization properties of ritonavir, fenofibrate, stability in solid dispersions is described in more
acetaminophen, and ABT-229 (a motilin receptor ag- detail in Section X.G.5.
onist structurally related to erythromycin) and also 2. Inhibition of Crystal Nucleation. The potential
showed little correlation between crystallization ten- for polymers to stabilize amorphous solid dispersions
dency and Tg. In this case, however, configurational via direct inhibition of nucleation is well described
entropy (and not enthalpy or free energy) and (Konno and Taylor, 2006; Marsac et al., 2006b; Caron
436 Williams et al.

et al., 2010). Thermodynamic drivers to nucleation depletes drug concentrations in the polymer in the
are explained by classic homogeneous nucleation vicinity of the crystal nucleus and drug-free polymer
theory (CNT) (Turnbull and Fisher, 1949; James, does not diffuse away from the interface. The local
1985), and concepts from the CNT have been applied concentration of polymer therefore increases, reducing
to solid dispersion systems to understand better the molecular mobility in the polymer-rich (high Tg) zone
stability of amorphous drug in the presence of and reducing drug access to the crystal nucleus.
polymers. From eq. 54, nucleation is also decreased by in-
The rate of nucleation (I) may be defined in terms of creasing DG* (i.e., the minimum free-energy change
the free energy of activation for molecular diffusion required to form a crystal nucleus) and, therefore, the
across the nucleation interface to join the new lattice thermodynamic barrier to crystal formation. As de-
(DGa) and the free-energy change associated with the scribed previously, DG* is dependent on DGs and DGv.
formation of a spherical nucleus of critical size (DG*) Increases in DG* can arise from an increase in DGs
(Konno and Taylor, 2006; Marsac et al., 2006a): after polymer accumulation at the nucleation interface
  and a change in interfacial tension between the two
2 ðDGa 1 DGp Þ phases, but such changes may also reflect changes to
I 5 A × exp  ð54Þ
kT DGv attributable to differences in free energy between
crystalline and amorphous drug. The latter is possible
where A is a constant, k is the Boltzmann constant, since drug-polymer interactions can lead to an increase
and T is absolute temperature. The free-energy change in the configurational entropy of the drug in the for-
(DG) during the formation of a spherical nucleus is mulation compared with unformulated drug in the
determined by the difference in the bulk free energy amorphous form (Konno and Taylor, 2006). Higher-
between the amorphous phase and the crystal phase configuration entropies may reduce the crystallization
(DGv) and the energy required to form a new solid tendency of amorphous drug since increased entropy
surface (DGs). DGv is volume dependent and favors increases the number of possible configurations that
nucleation as the free energy of the crystal is lower the drug can hold, decreasing the probability that drug
than that of the amorphous form (i.e., DGv is negative). molecules are correctly orientated for crystallization to
DGs is a function of the surface area of the new phase occur (Zhou et al., 2002; Marsac et al., 2006a; Baird
and interfacial tension and disfavors nucleation as and Taylor, 2011). However, relationships between the
energy is required to form a new surface (i.e., DGs is crystallization of amorphous drug and entropy are not
positive). As the size of the crystal nucleus increases, always apparent (Konno and Taylor, 2006; Marsac
the magnitude of DGv increases more rapidly than DGs et al., 2006a; Bhugra and Pikal, 2008).
(since volume increases more rapidly than area with 3. Molecular Mobility, Strength, and Fragility.
increasing diameter). DG therefore reaches a maxima Molecular mobility and associated descriptors of
in the DG versus crystal nucleus size plot, after which mobility such as relaxation, fragility, and strength
DG decreases without limit. This maxima is the free are commonly used to describe the physical stability of
energy for formation of a nucleus of critical size (DG*), amorphous materials and solid dispersion formula-
and it defines the mimium (critical) nucleus size for tions. A degree of molecular mobility is required to
crystallization. Before reaching the critical nucleus effect nucleation and crystallization; hence, decreases
size, small nuclei may still form but require additional in mobility are expected to reduce crystallization rates
free energy for growth. As such, their existence is only (Hancock et al., 1995; Shamblin et al., 1999, 2000).
transient, and the nuclei rapidly decompose. However, Molecular mobility can be reduced by increasing vis-
above the critical nucleus size, crystal growth is cosity, and the decrease in mobility inherent in an in-
thermodynamically favored. crease in Tg has been implicated in the potential utility
DGa is often incorporated into the pre-exponential of many excipients to enhance the stability of solid
term in eq. 54 (James, 1985; Bhugra and Pikal, 2008) dispersions (Van den Mooter et al., 1998, 2001; Verreck
but can also be expanded to enable discussion of the et al., 2003b; Sethia and Squillante, 2004; Six et al.,
potental impact of polymers on nucleation rates in 2004). Mobility is also reduced by intermolecular
solid dispersions (Konno and Taylor, 2006; Marsac interactions between drug and excipient (polymer)
et al., 2006a). Konno and Taylor (2006) suggest that (Taylor and Zografi, 1997; Matsumoto and Zografi,
polymers increase DGa (and, therefore, reduce nucle- 1999; Khougaz and Clas, 2000; Gupta et al., 2004;
ation rate in eq. 54) by three potential mechanisms: 1) Rawlinson et al., 2007).
by acting as a diluent and providing a physical im- During storage, the molecular structure of a thermo-
pediment to nucleation; 2) by allowing for specific in- dynamically unstable amorphous material relaxes and
teractions between the drug and polymer, reducing the degree of relaxation on storage may be used as an
drug diffusivity; and 3) by accumulating at the nu- indicator of molecular mobility and stability. Relaxation
cleation interface. Accumulation of a polymer-rich layer and diffusion in glassy materials are complex and
at the nucleation interface occurs when drug crystallization multifactorial and have been widely studied using
Strategies for Low Drug Solubility 437

experimental and theoretical approaches, especially in subsequently related to the degree of relaxation, which
the past decade (Capaccioli et al., 2008). Simplistically, in turn is a function of molecular mobility: the greater
relaxation may be subdivided into primary a-relaxation the molecular mobility, the greater the degree of
and secondary b-relaxation (Priestley et al., 2005). relaxation.
a-Relaxation involves rotational and translation re- The enthalpy loss on relaxation (DHrelax) is esti-
orientation of molecules in a global sense (i.e., through- mated via measurement of the enthalpy recovered
out a sample) and is most apparent at temperatures during sample reheating past the Tg (Hancock et al.,
above and immediately below the glass transition 1995) and compared with DH‘ (the maximum amount
temperature (i.e., where there is greater mobility). of recovered enthalpy at infinite time) to give the
b-Relaxations are observed at lower temperatures and extent of relaxation (F) from eq. 55:
include localized rotation of molecules or molecular  
segments; b-relaxation processes include the well- DHrelax
F512 ð55Þ
described Johari-Goldstein (J-G) b-relaxations (Johari DH‘
et al., 2005; Prevosto et al., 2009), which followed the
work of Johari and Goldstein (1970), who identified DH‘ is calculated from the change in heat capacity (DCp)
secondary relaxation behaviors within rigid glass struc- of the drug between the glass transition temperature
tures (i.e., below Tg). Both a- and b-relaxations appear (Tg) and the annealing temperature (T):
to occur universally in glass-forming materials (Johari  
and Goldstein, 1970; Thayyil et al., 2008). Evidence DH‘ 5 DCp Tg 2 T ð56Þ
that a-relaxation (as captured by methods describing
relaxation and mobility changes on movement through Where repeated measurements are obtained, the data
Tg) cannot always be linked to changes in the physical describing relaxation (F) as a function of the annealing
state of amorphous solids (Johari et al., 2005; Vyazovkin (storage) time (t) can be compared with models such as
and Dranca, 2005) and a realization that J-G b- the Kohlrausch-Williams-Watts (KWW) equation that
relaxations are precursors of longer-range motion, describe the approach to a fully relaxed state:
including a-relaxation and the glass transition (Ngai,   b
T
1998; Johari et al., 2005; Prevosto et al., 2009), suggest F 5 exp 2 ð57Þ
that J-G b-mobility may better describe the molecular t
changes that lead to nucleation (Vyazovkin and Dranca, This provides an average relaxation time constant t
2005; Bhugra and Pikal, 2008). Differences in J-G and a stretch parameter b that varies between 0 and 1
b-mobility below Tg may therefore provide an improved to indicate deviation from exponential relaxation
indication of crystallization potential in amorphous behavior. Deviations are often seen since the level of
solids (Ngai et al., 2001; Tombari et al., 2008a,b). The structural relaxation during sample annealing may
contribution of a- and b-relaxation to stability in amor- vary considerably within a sample and during the
phous pharmaceuticals is well described elsewhere course of the annealing experiment. Relaxation is
(Bhattacharya and Suryanarayanan, 2009). therefore nearly always nonlinear and nonexponential
Molecular mobility can be estimated via measurement (Davies and Jones, 1953).
of viscosity, dielectric relaxation time, or structural Structural relaxation is highly temperature depen-
relaxation time, and a range of techniques—including dent, and trends in relaxation as a function of temp-
rheometry, dielectric, spectroscopy, and solid-state erature have been used to provide additional insight into
NMR—have been used to quantify these properties the potential stability of amorphous materials. Several
(Andronis and Zografi, 1997; Yoshioka et al., 1999; Aso models have been used to describe the temperature
et al., 2000; Liu et al., 2002; Offerdahl et al., 2005). In dependence of relaxation time. The Vogel-Fulcher-
pharmaceutical applications, most studies have used Tammann (VFT) equation (eq. 58) is commonly used to
thermal methods, in particular DSC, to provide an describe relaxation above Tg, whereas the Adam-Gibbs-
indication of changes to molecular mobility. As de- Vogel (AGV) equation (eq. 59) is used below the Tg:
scribed in the following section, this approach has
 
limitations, but widespread application warrants a de- DT0
t 5 t0 exp ð58Þ
scription of the methods used. In general, thermal T 2 T0
analytical approaches to estimate molecular mobility
0 1
within a glassy solid (below the Tg) are performed
using annealing experiments or “enthalpy recovery B DT0 C
t 5 t0 expB
@  C ð59Þ
experiments.” These are described in detail elsewhere T0 A
(Hancock et al., 1995; Mao et al., 2006a) but rely on T 12
Tf
molecular relaxation during storage below Tg (anneal-
ing), resulting in a loss of enthalpy. The loss of In eq. 58 and eq. 59, T0 is the temperature at which
enthalpy (i.e., the relaxation enthalpy, DHrelax) is the configurational entropy is zero (i.e., no further
438 Williams et al.

relaxation is possible) and has the same physical susceptible to crystallization (Fig. 58). The data also
meaning as the Kauzmann temperature (Fig. 56), confirmed the importance of drug load in determining
t 0 is the relaxation time constant for an unrestricted physical stability (Friesen et al., 2008), and at higher
material and is normally taken as being of the order of drug loads, the stabilizing effect of the polymer de-
the lifetime of atomic vibrations (10214 seconds) (Mao creased. For poor glass-formers (i.e., those with higher
et al., 2006a), and Tf is the fictive temperature (the Tm:Tg ratios in Fig. 58), adequate stability was
temperature at which the structure of a supercooled achieved only at low drug loads (i.e., where the
liquid is “frozen-in” and quantitatively similar to the polymer-to-drug ratio was high). The importance of
Tg) (Badrinarayanan et al., 2007). Tf is determined by drug loading in solid dispersion stability is described in
the point of intersection of lines extrapolated from more detail in Section X.G.5.
above and below the Tg in thermal melting plots Given the link between drug fragility and the long-
(similar to Fig. 56), and errors in Tf estimation may term stability of formulated solid dispersions, attempts
lead to significant differences in shelf-life estimations have subsequently been made to understand better the
(Johari and Shanker, 2010). The strength parameter molecular properties that predetermine fragility. Us-
(D) provides an alternate measure of the fragility of ing principal component analysis, poor glass-formers,
pharmaceutical glasses. Most drugs are classified as that is, those showing high crystallization tendencies
either fragile (D , 10) or moderately fragile (D 5 and lower physical stability (defined as “class I”
10–30) (Crowley and Zografi, 2001; Zhou et al., 2002; molecules), were shown to have lower molecular
Mao et al., 2006a), and for these materials, additional weights and fewer rotational bonds than compounds
formulation approaches are typically needed to facili- showing higher glass forming potential (class III
tate glass formation and to ensure stability. Significant molecules) (Baird et al., 2010). This was suggested to
variations in crystallization tendency also exist even reflect differences in configurational entropy, such that
within the fragile or moderately fragile classifications a reduction in the number of rotational bonds (a
(Baird et al., 2010). As described in the preceding situation that is more likely for drugs with lower
section, annealing experiments are widely used to molecular weights) increased the probability of the
measure the level of molecular mobility below glass drug existing in the correct orientation for crystalliza-
transition temperatures (Andronis and Zografi, 1998; tion (Baird et al., 2010). Class II compounds showed
Shamblin et al., 1999; Liu et al., 2002; Weuts et al., a similar level of glass-forming potential to class III but
2003; Mao et al., 2006a,b) and were used, for example,
by Zografi and coworkers (Hancock and Zografi, 1997)
to establish the Tg–50°C rule, described in Section X.
F.3.
Tm:Tg or Tg:Tm ratios (the latter ratio being the
reduced glass transition temperature, Trg) have also
been used as an alternative indicator of fragility (and
therefore stability or glass forming ability) where

Tg
Trg 5 ð60Þ
Tm
Most drugs have Trg values between 0.65 and 0.8
(Baird and Taylor, 2012) and the approximation that
Trg 5 0.66 is widely used to predict glass transition
temperatures from the melting temperature (the so-
called “2/3 rule”). Poor glass-formers, however, show
significantly higher melting temperatures relative to
their glass transition temperatures and exhibit lower Fig. 58. Tm/Tg versus log P for low-solubility compounds successfully
Trg, (typically Trg , 0.66) with decreasing Tg:Tm ratios formulated as solid dispersions (all achieved supersaturation during in
vitro dissolution tests). The boundaries of regions (i) to (iv) are arbitrary
indicative of materials with greater fragility and but are representative of general trends in decreasing physical stability
reduced glass-forming ability. with decreasing Tg of the drugs [moving vertically from region (i) up to
region (iii)] and decreasing dissolution rate with increasing drug
Friesen et al. (2008) investigated the ability of the lipophilicity [moving horizontally from region (i), (ii) or (iii) into region
reduced glass transition temperature to estimate the (iv)]. Because of the high Tg values, drugs in region (i) may be
potential for polymers to enhance the stability of solid incorporated into solid dispersions at higher (.50% w/w) drug loadings;
the lower stability of drug in the amorphous form in regions (ii) and (iii)
dispersion formulations. This study examined the pro- requires lower drug loading for long-term solid dispersion stability
perties of HPMCAS spray-dried solid dispersions for (35–50% w/w and 10–35% w/w, respectively). Drug loading in region (iv) is
low (10–35% w/w) and limited by a slow rate of dissolution of lipophilic
139 experimental compounds. Across the data set, drugs. Tm, melting temperature; Tg, glass transition temperature.
drugs with higher Tm:Tg ratios (lower Trg) were more Adapted from Friesen et al. (2008).
Strategies for Low Drug Solubility 439

greater crystallization tendencies—a property attrib- that of the solid dispersion). The advantages and
uted to the ease of formation of crystal nuclei during limitations of various instrumental techniques to
cooling of drug melts. Class I compounds (i.e., those measure molecular mobility are discussed further in
least likely to form amorphous material) included Bhattacharya and Suryanarayanan (2009).
tolfenamic acid, griseofulvin, carbamazepine, and tol- 5. Drug-Polymer Miscibility and Relationship to Drug
butamide. Class II included acetaminophen, celecoxib, Loading. Solid dispersions containing drug molecu-
nifedipine, and cinnarizine; class III (i.e., those with larly dispersed throughout the polymer matrix (i.e.,
a higher glass-forming potential and most stable in the types 3 and 6 systems as defined in Table 26) are
amorphous form) included fenofibrate, felodipine, itra- typically more resistant to drug crystallization. In
conazole, ritonavir, and indomethacin (Baird et al., contrast, systems containing higher local concentra-
2010). Mahlin et al. (2011) used in silico techniques to tions of phase separated amorphous drug (i.e., types 2
predict glass-forming potential based on molecular and 5) increase the likelihood of molecular alignment
structure, and consistent with the data described by and nucleation and are more prone to crystallization.
Baird et al. (2010), suggest that larger molecules with Physical instability in solid dispersions is therefore
larger numbers of rotational bonds and limited molec- more common where there is heterogeneous mixing,
ular symmetry show good glass-forming potential. where drug-polymer miscibility is poor and where
4. Limitations to the Use of Thermal Analysis to drug loading exceeds the drug-polymer miscibility
Estimate Molecular Mobility. Although estimations limit (van Drooge et al., 2006; Marsac et al., 2009;
of molecular relaxation using changes in heat capacity Paudel et al., 2010; Qian et al., 2010a). Understand-
remain common, in large part because of the wide ing the relationship between drug loading and the
availability of thermal analytical techniques, Johari drug-polymer miscibility limit (and therefore the
and Shanker (2010) have highlighted a number of molecular arrangement of drug in the formulation)
potential shortcomings of correlating measured or is critical to the development of amorphous solid dis-
calculated thermodynamic quantities (such as heat persions that maintain physical stability over pro-
capacity of a glass) with the molecular mobility of an longed periods.
amorphous pharmaceutical to estimate differences in The miscibility of amorphous drug in a polymer
stability. In these studies, relaxation dynamics in matrix is dependent only on adhesive and cohesive
acetaminophen glasses were instead obtained using interactions between drug and polymer (whereas the
dielectic relaxation spectroscopy (Johari et al., 2005) solubility of crystalline drug is also dependent on the
and the data compared with previous studies using strength of the crystal lattice). For amorphous drug,
calorimetry (Zhou et al., 2002). In short, the data where drug-polymer miscibility is thermodynamically
obtained using thermal analysis appeared to underes- favored, the free energy of mixing (DGmix) is negative
timate relaxation. The authors suggest that this (eq. 61):
finding reflected errors in estimation of configurational DGmix 5 DHmix 2 TDSmix ð61Þ
entropy from specific heat data since entropy originat-
ing from nonconfigurational sources may contribute The mixing of two components typically increases
significantly to heat-capacity changes on varying entropy (DSmix), favoring miscibility. Mixing enthalpy
temperature but does not appear to have a role in (DHmix) is therefore the more discriminating factor in
structural relaxation (Johari et al., 2005). Dielectric determining miscibility, with negative or slightly posi-
relaxation spectroscopy may therefore represent a more tive values favoring spontaneous miscibility (Pajula
accurate means of assessment of molecular mobility, et al., 2010). The value of DHmix is a function of the
although some limitations for pharmaceutical applica- relative strength of cohesive interactions (drug-drug
tions are apparent (Craig, 1992). Specifically, in their and polymer-polymer interactions) and adhesive inter-
most preferred configuration, dielectric spectroscopy actions (drug-polymer interactions), where negative
measurements are conducted using a sealed cell in mixing enthalpy (and therefore miscibility) is favored
which the sample is melted (Alie et al., 2004; Johari when adhesive interactions dominate.
et al., 2005; Adrjanowicz et al., 2010a; Dantuluri et al., A single Tg value during a melting DSC cycle is
2011). In contrast, pharmaceutical scientists are usually considered satisfactory evidence of drug-
usually interested in monitoring stability over the polymer miscibility (assuming the Tg of the drug and
projected pharmaceutical shelf-life and in samples that polymer are different). In contrast, multiple Tg values
have been subjected to different humidity and temper- suggest regions of heterogeneity resulting from in-
ature conditions. This requires the use of a “parallel complete drug-polymer mixing or limited drug-polymer
plate” method (Adrjanowicz et al., 2010a; Grzybowska miscibility (Six et al., 2003a,b; van Drooge et al., 2006).
et al., 2010); however, this may introduce additional DSC is therefore routinely used to determine the
experimental error, including the entrapment of miscibility of drug-polymer mixtures. However, drug-
varying amounts of air and moisture (the dielectric polymer miscibility at elevated temperatures (i.e.,
constants of air and water are markedly different from temperatures around the Tg) is not necessarily
440 Williams et al.

predictive of miscibility at lower (storage) temper- Higher-resolution imaging of drug distribution


atures (Qi and Craig, 2010; Qian et al., 2010a,b; Baird patterns in solid dispersions may be achieved using
and Taylor, 2012). Similarly, the lack of crystalline tapping-mode atomic force microscopy and has been
peaks in x-ray diffraction data (i.e., the presence of the used to assess a range of drug-polymer films and
characteristic amorphous halo) provides evidence of powders for areas of immiscibility (with resolution
amorphous content but little additional information as down to the nanometer range) (Weuts et al., 2011).
to miscibility of the amorphous drug with the polymer Example images from a recent study are reproduced
(i.e., the degree of molecular dispersion). in Fig. 60 (Lauer et al., 2011). In this case, the drugs
The limitations of DSC and XRPD in assessing solid [a neurokinin 1 (NK1) receptor antagonist and an
dispersion physical stability were recently highlighted inhibitor of cholesterylester transfer protein (CETP)]
by Qian et al. (2010b), who used hot-melt extrusion to show evidence of miscibility with PVP, PVPVA, and
prepare PVPVA solid dispersions of an unnamed ex- HPMCAS (i.e., topography consistent with that of the
perimental compound, using either a low (50 rpm) or pure polymer), whereas discrete particles (;80 nm)
high (225 rpm) extruder screw speed. On initial as- are observed in Eudragit L100 films, suggesting
sessment, batches produced using either method had phase-separation (Lauer et al., 2011). Friesen et al.
a single and equivalent Tg (88°C). XRPD assessments (2008) also suggest that scanning electron micros-
were also similar, and both methods of characterization copy can be used to detect crystallinity in solid
suggested that variation in screw speed had little effect dispersions at levels as low as 1%. Further details of
on the properties of the solid dispersion. In contrast, the techniques that can be used to detect phase-
repeat testing after 2 months revealed that crystalliza- separated drug within solid dispersions are provided
tion had occurred in the batch produced using the slower in two recent reviews (Qi and Craig, 2010; Baird and
screw speed, whereas the batch prepared using the high Taylor, 2012).
screw speed was stable. Confocal Raman microscopy To circumvent some aspects of the empirical nature
was subsequently used to chemically map samples of the of solid dispersion design and to inform the selection of
two solid dispersions. Significant differences in homoge- carriers that are miscible with the drug of interest,
neity were obtained (Fig. 59), with the least stable solid initial estimates of drug-polymer miscibility may be
dispersion (left) showing regions rich in amorphous drug achieved using the Flory-Huggins theory of mixing of
(red areas in Fig. 59). These data provide a good regular solutions (Lloyd et al., 1997; Yao et al., 2005;
example of the potential advantage of characterization Marsac et al., 2006b, 2009; Pajula et al., 2010; Sun
methods beyond conventional DSC and XRPD, in et al., 2010; Baird and Taylor, 2011) or Hildebrand
particular, spatially specific spectroscopic methods, to solubility parameters (Greenhalgh et al., 1999; Forster
assess the homogeneity of mixing in amorphous solid et al., 2001b; Paudel et al., 2010; Huang et al., 2011;
dispersions (Breitenbach et al., 1999; Qian et al., 2010b). Zhao et al., 2011).

Fig. 59. Representative compositional Raman images of two experimental solid dispersions prepared by hot melt extrusion (HME 1 and HME 2,
respectively). The red regions depict higher local concentration of drug in the formulation (an undisclosed compound). The size of each Raman map is
800 mm  800 mm and 1600 (40  40). Raman spectra were collected to construct each image. HME 1 5 slow extruder screw speed (50 rpm); HME 2 5
fast extruder screw speed (225 rpm). Reprinted from Qian F, Huang J, Zhu Q, Haddadin R, Gawel J, Garmise R, and Hussain M (2010b) Is a distinctive
single Tg a reliable indicator for the homogeneity of amorphous solid dispersion? Int J Pharm 395:232–235, with permission from Elsevier.
Strategies for Low Drug Solubility 441

The Flory-Huggins equation for mixing of two poly- between molecules of a single component are stronger
mers (eq. 62) (Flory, 1953) is shown below and is an than those between unlike pairs, causing the third
extension of the Gibbs free-energy change on mixing term in eq. 62 to oppose mixing. This term is governed
(eq. 61): by the value of the Flory-Huggins interaction param-
eter (x). A large positive x value causes the enthalpy
    component to dominate the equation and promotes
fΑ fB
DGmix 5 RT lnfA 1 lnfB 1 x  fA   fB ð62Þ immiscibility of the system. In contrast, a negative or
rA rB
marginally positive value for x suggests good mis-
In the Flory-Huggins equation, R and T are the gas cibility (Paudel et al., 2010).
constant and temperature, fA and fB are volume The dependence of DGmix on the quantity of drug in
fractions of the two polymer components being mixed, a solid dispersion is evident by plotting the solution of
rA and rB are the number of monomer segments in each the Flory-Huggins equation (eq. 62) as a function of the
polymer component, and x is the interaction parameter volume fraction of drug (fA). At fixed temperature, this
between these two components. In the case of a solid plot has only one independent variable (fA). However,
dispersion, A and B refer to drug and polymer com- the shape of the curve is temperature dependent as x
ponents of the mixture and for the drug, rA 5 1. varies inversely with temperature. At low x values
The first two terms on the right in eq. 62 represent (i.e., high temperature), all compositions are miscible,
the contribution of the combinatorial entropy of mixing whereas at high x values, immiscibility results. Most
of the two components, while the third term represents combinations of polymer and drug are not miscible over
all noncombinatorial factors included in the enthalpy all temperature ranges. For these partially miscible
of mixing. Since mixing leads to an increase in entropy, systems, a plot of DGmix versus fA shows a central
the first two terms in the Flory-Huggins equation maxima flanked by two free-energy minima. An
almost always favor mixing. The third term represents example of such a plot is shown schematically in Fig.
the change in local interactions and motions of the 61A. At a fixed temperature, miscibility is favored by
drug and polymer. In most cases, attractive energies volume fractions close to the free-energy minima. In

Fig. 60. Nanometer homogeneity in various drug-polymer mixtures is shown with phase maps recorded with tapping-mode atomic-force microscopy on
fracture surfaces of the blank excipients (upper panels, A–E) and corresponding combinations with a neurokinin 1 (NK1) receptor antagonist (middle
panels, F–J) and also with an inhibitor of cholesterylester transfer protein (CETP) (bottom panels, K–O). The numbers visible in the image rows (F–J
and K–O) refer to the number of imaged regions across the sample showing comparable surface morphology to that illustrated in the images above
relative to the total number of regions imaged. The scale bar is 1 mm in length. Reprinted from Lauer et al. (2011), Copyright 2011, with permission
from Springer.
442 Williams et al.

solid dispersions, phase separation in this way is


hindered by the low kinetic mobility of the dispersion
(which limits diffusion of drug molecules within the
polymer matrix) and the energy required to form
a drug nucleus of critical size. The importance of local
molecular mobility in dictating nucleation and growth
is consistent with an important role for secondary or
b-relaxation behavior in the physical stability of solid
dispersions below Tg. In contrast, for volume-fraction
compositions and temperatures located inside the
spinodal line (the unstable region), the system is highly
unstable to even small molecular fluctuations, and
phase separation takes place spontaneously throughout
the sample via spinodal decomposition (rather than
local nucleation) and only diffusion limits phase se-
paration. The global nature of spinodal decomposition
suggests a closer link to primary or a-relaxation be-
havior. Knowledge of whether the composition of a
solid dispersion exists between the phase boundary and
spinodal curve (metastable region) or inside the spino-
dal curve (unstable region) at the storage temperature
provides an additional indication of physical stability
(Fig. 61B). Although both systems are thermodynam-
Fig. 61. Schematic depicting (A) the effect of increasing fA (the volume ically unstable, the risk of spontaneous phase sepa-
fraction of component (A) on the free energy of mixing (ΔGmix) at ration via spinodal decomposition (in the case of
a constant temperature, (B) miscibility boundaries as a function of systems that lie inside the spinodal curve) typically
temperature and increasing fA. The green curve in (A) is calculated using
the Flory-Huggins interaction parameter x and eq. 62. The free-energy precludes the preparation of solid dispersions with
minima are highlighted by the letter x. Taking the first derivative of eq. acceptable physical stability over prolonged periods.
62 and setting ΔGmix equal to zero allow the phase boundary curve in (B)
to be derived, and by taking the second derivative of eq. 62, it is possible In contrast, appropriate polymer choice may allow
to derive the spinodal curve. The second derivative is equal to zero at the kinetic stabilization of metastable compositions and
spinodal points on the green curve in (A) (i.e., the letter y). Mixtures that
lie above the phase boundary are miscible and stable. Mixtures that lie
facilitate exploitation of formulations with higher
inside the phase boundary line but outside the spinodal curve are drug load.
metastable. Mixtures that lie inside the spinodal curve (the spinodal Solubility parameters (Hildebrand and Hansen) pro-
region) are unstable.
vide an alternative mechanism for estimation of the
potential for drug-polymer miscibility, where drug-
polymer combinations with the most similar solubility
the central range of compositions, the maxima in the parameters are expected to provide the greatest likeli-
free energy profile leads to immiscibility. Since this hood of miscibility. Huang et al. (2011) examined the
immiscible portion of the curve exists between two miscibility of polymeric solid dispersions of nifedipine
miscible regions, it represents a miscibility gap. and showed that predictions of miscibility using the
Once x has been determined from the Flory-Huggins Flory-Huggins parameter and Hildebrand solubility
equation (eq. 62) and the dependence of the volume parameters both indicated greater nifedipine miscibil-
fraction of drug on DGmix has been plotted (i.e., Fig. ity in Eudragit RL compared with ethylcellulose. These
61A), a spinodal curve can be calculated by taking the predictions were subsequently verified experimentally
second derivative of eq. 62. The spinodal curve defines using FTIR and DSC. Similarly, Paudel et al. (2010)
(and separates) the metastable and unstable regions examined naproxen-PVP systems and showed that the
within the two-phase (i.e., nonmiscible) portion of Flory-Huggins interaction parameter was large and
the phase diagram (shown in Fig. 61B) (Favvas and negative, and that this correlated well with high
Mitropoulos, 2008). Compositions within these two kinetic miscibility limits (. 50% w/w drug in PVP)
regions phase-separate via different mechanisms. Meta- for solid dispersions prepared by spray drying. In con-
stable compositions are located between the phase trast, however, when the solubilities of crystalline and
boundary and the spinodal curve and are relatively amorphous naproxen in PVP were estimated using
stable to small molecular fluctuations. Phase separation solubility in a low-molecular-weight PVP analog (n-
requires a significant local perturbation to the system methylpyrrolidone), solubility was low (5–7% w/w) in
and can be achieved only by the formation of a nucleus both cases, suggesting that PVP solid dispersions
of critical size and subsequent particle growth (Favvas prepared at drug loadings greater than ;7% w/w were
and Mitropoulos, 2008). In the case of drug-polymer highly supersaturated and, despite Flory-Huggins
Strategies for Low Drug Solubility 443

predictions of high miscibility, at risk of phase se- Figure 62 provides a simplified interpretation of the
paration. The limitations of the Flory-Huggins model relationships between drug loading relative to the
in this example were ascribed to an inability to account crystalline solubility/amorphous miscibility limits in
for specific saturable interactions (e.g., drug-polymer a solid dispersion and physical stability. Thermody-
H-bonds) and the fact that the interaction parameter namically stable solid solutions are formed where drug
was based on mixing data at high temperatures (i.e., load is lower than the crystalline drug solubility in the
near the melting temperature of the drug) where mis- carrier (assuming homogeneous dispersion of drug
cibility was favored (eq. 62) (Paudel et al., 2010). within the carrier) (Vasanthavada et al., 2004; Qian
Since attractive intermolecular interactions between et al., 2010a). However, this situation is rare since
drug and polymer provide a driver for enhanced mis- crystalline solubility is invariably low and below the
cibility, several spectroscopic techniques, including in- concentrations required to provide an appropriate dose.
frared, Raman, and NMR spectroscopy, have been used For example, the crystalline solubility of itraconazole
to probe the presence of drug-polymer interactions in in Eudragit E100 is 0.012% w/w, whereas the amor-
solid dispersions (Taylor and Zografi, 1997; Van den phous solubility (i.e., miscibility) is approximately 7%
Mooter et al., 1998; Tantishaiyakul et al., 1999; Khougaz (Janssens et al., 2010).
and Clas, 2000; Konno and Taylor, 2006; Marsac et al., Higher drug loads (i.e., those greater than the crys-
2006a; Rumondor et al., 2009a). The absence of drug talline solubility limit) result in solid solutions that are
polymer interactions, however, does not always pre- supersaturated with respect to drug crystal solubility.
clude miscibility, and Eudragit E100 and itraconazole, These systems are thermodynamically unstable, and
for example, are partially miscible but show no evi- over time, the drug is susceptible to crystallization.
dence of intermolecular bonding (Janssens et al., 2010). Where drug crystallization occurs, a two-phase solid

Fig. 62. Scheme depicting a simplified version of events that occur on increasing drug loading in solid dispersions in relation to physical stability.
Individual drug molecules are depicted by the yellow triangles and are present in a carrier matrix (green). Increasing drug concentrations in the solid
dispersion provides additional utility through higher potential drug doses but is accompanied by additional stability concerns. Solid dispersions
containing drug above the crystalline solubility are thermodynamically unstable and may separate to form multiphase systems via nucleation and
particle growth. Above the amorphous solubility (miscibility limit), solid dispersions initially phase separate by two different mechanisms. Within the
metastable region (i.e., between the miscibility curve and the spinodal curve; see also Figure 61B), separation occurs via nucleation and growth.
However, in the unstable region (i.e., below the spinodal curve), solid dispersions spontaneously phase-separate. In the event of spontaneous phase
separation, crystallization is expected to occur rapidly as concentrations of stabilizing polymer surrounding the drug molecules are low. The positions of
the solubility limits (crystalline and amorphous) are carrier dependent for a particular drug, and phase separation may occur below the amorphous
solubility (miscibility limit) if heterogeneous mixtures are prepared during manufacture or during storage.
444 Williams et al.

dispersion is produced comprising crystalline drug more likely to crystallize compared with solid disper-
particles suspended in the carrier (i.e., reversion to sions where drug is present in a molecularly dispersed
type 1 or type 4 solid dispersions dependent on the form (i.e., type 3 or type 6 systems) that has increased
physical form of the carrier). The potential for crys- kinetic stability as a result of intimate contact with the
tallization to occur via the initial formation of amor- polymer (Six et al., 2004; Marsac et al., 2006b;
phous nuclei or particles (i.e., type 2 or type 5 systems) Rumondor and Taylor, 2010). Solid dispersions with
is also shown in Fig. 62, although this intermediate lower drug-polymer ratios and those with high drug-
may be transient, and spontaneous crystallization of polymer miscibility therefore show better physical
drug from a supersaturated solid solution is also stability at ambient temperatures (Chan and Kazarian,
possible (Rumondor and Taylor, 2010). 2004; Van Drooge et al., 2004; Weuts et al., 2005;
The greater the drug loading relative to the crys- Friesen et al., 2008; Qian et al., 2010a). These solid
talline solubility, the higher the level of supersatura- dispersions are also more physically stable at elevated
tion and the higher the driving force for crystallization temperature and in the presence of moisture (Van
[since the thermodynamic barriers to crystallization Drooge et al., 2004; Vasanthavada et al., 2004;
(i.e., ΔG*) are reduced (eq. 54)]. Increasing drug Rumondor and Taylor, 2010). Water or other solvents
loading ultimately exceeds the amorphous drug solu- (that may be carried over from manufacturing) reduce
bility in the carrier (i.e., the drug-polymer miscibility miscibility by promoting molecular interactions be-
limit). Above the miscibility limit, physical instability tween solvent and drug or polymer and, therefore,
occurs via two separate mechanisms. Where the drug- reduce the number of possible drug-polymer intermo-
polymer composition is located between the phase lecular interactions. The plasticizing properties of sol-
boundary and spinodal curve (see preceding discus- vents also encourage phase-separation in situations
sion), solid dispersions are metastable and phase- where drug loads exceed the miscibility limit through
separate via nucleation and particle growth. This increased molecular mobility in the system. Rumondor
generates areas of localized high concentration of and Taylor (2010) further propose that moisture affects
amorphous drug in a manner analogous to that seen the route of drug recrystallization, either by promoting
below the amorphous solubility limit. At higher drug phase-separation, and therefore crystallization from
loads, beyond the spinodal curve, solid dispersions are heterogeneous dispersions containing amorphous drug
unstable and spontaneously phase-separate by spinodal clusters, or by promoting spontaneous crystallization
decomposition rapidly to form domains rich in amor- from homogeneous dispersions (see Fig. 62). The latter
phous drug. The localized concentrations of polymer route of crystallization and crystal growth is likely to
surrounding areas of phase separated amorphous drug be slower given the higher surrounding concentrations
are low, increasing the risk of drug crystallization (Six of stabilizing polymer.
et al., 2004; Vasanthavada et al., 2004; Rumondor and Finally, the processes depicted in Fig. 62 assume
Taylor, 2010). initial preparation of homogeneous solid dispersions.
Changes in physical form of a drug within solid Where the method of manufacture does not provide for
dispersions with changing drug load may be monitored adequate mixing, drug and polymer phase-separation
via effects on Tg in situations where the drug plas- may occur below the miscibility limit. Under conditions
ticises the polymer (i.e., reduces polymer Tg) (Six et al., of poor mixing, regions exist that are drug rich and
2004; Van Drooge et al., 2004, 2006; Vasanthavada polymer poor, and the (local) level of drug in these
et al., 2004). For example, in diazepam-PVP solid regions is likely to be in excess of the miscibility limit.
dispersions (prepared by freeze-drying), increasing This scenario leads to localized phase separation and
drug load leads to a progressive decrease in the Tg of drug crystallization, which may ultimately trigger
the drug-polymer mixtures, and above 35% w/w drug, more widespread (global) crystallization.
two distinct Tg emerge (i.e., drug and polymer glass
transitions) as the drug-phase separates from the H. Summary
polymer. At higher drug loads, the absolute level of Solid dispersion formulations comprise drug physi-
molecularly dispersed drug is largely unchanged, cally dispersed within an inert and usually highly
however, the level of drug arranged as amorphous water-soluble carrier, often a high molecular weight
clusters increases steadily. Solid dispersions greater polymer. Drug may exist in the carrier in multiple
than the diazepam-PVP miscibility limit (35%), there- physical forms and may be dissolved, partially dis-
fore, show characteristics of both an amorphous solved with excess drug suspended in the crystalline
suspension (type 5) and an amorphous solid solution state, partially dissolved with excess drug in the
(type 6) (van Drooge et al., 2006). amorphous state, or partially dissolved with excess
Unlike molecularly dispersed formulations, phase- drug present as a mixture of crystalline and amor-
separated drug present as suspended amorphous phous states. The inert carrier matrix may also be
particles is no longer intimately mixed with the carrier. present in either crystalline or amorphous forms. Aca-
This more heterogeneous molecular rearrangement is demic researchers and industrial investigators have
Strategies for Low Drug Solubility 445

used myriad terms to describe “solid dispersions” that stability below Tg. A deep understanding of the drivers
comprise combinations of these physical states of of recrystallization, however, remains elusive, and the
solute and matrix. These include solid solutions, use of thermal methods to estimate thermodynamic
eutectic mixtures, solid amorphous dispersions, molec- indicators of stability may incompletely capture the
ular dispersions, amorphous molecular dispersions, complexity of what is essentially a kinetic process.
coprecipitates, and sugar glasses. All are used exten- Ongoing studies in material sciences and physics to
sively in the pharmaceutical literature without the understand and predict better the behavior of amor-
existence of a standard or widely accepted nomencla- phous and glassy solids should lead to better un-
ture. Here we simplify to account for six classes of solid derstanding of physical stability and ultimately in
dispersion where drug is present either molecularly vitro/in vivo performance of solid dispersions. The
mixed in solution or as a crystalline or amorphous continuing challenge for pharmaceutical scientists is to
dispersion and where the carrier is amorphous or keep abreast with the advances in this area and to use
crystalline (Table 26). Of these, systems containing these studies to build confidence and robustness into
amorphous drug provide for the most significant the development of drug products containing amor-
increases in solubility, dissolution, and oral bioavail- phous solids.
ability and are therefore most widely used. Indeed, the
use of noncrystalline solids as a single or multicompo-
XI. Lipid-Based Formulations
nent system may be one of the most efficient means of
achieving enhanced drug solubilization, especially for The physicochemical properties that predispose
drugs with relatively high melting points. poorly water-soluble drugs to low and variable oral
Despite the great potential of solid dispersions, bioavailability (i.e., low water solubility, high log D)
especially amorphous solid dispersions, for bioavail- are shared by many dietary lipids, fat-soluble vitamins,
ability enhancement, there are few successful exam- and sterols. Dietary lipids and lipophilic nutrients,
ples of products on the market. This reflects a great however, are typically well absorbed, even after in-
many complexities with their use, in particular the gestion of lipid “doses,” which can be as high as 100 g
intrinsic thermodynamic instability of drug in the per day, as part of an adult’s Western diet. The effi-
noncrystalline form. High molecular-weight polymers ciency of lipid absorption reflects the evolution of
have a proven capacity to kinetically stabilize drug in highly specialized lipid-processing pathways that cir-
the high-energy amorphous form, however, many fac- cumvent the problems of low water solubility and lead
tors contribute to appropriate physical stability. These to efficient solubilization of dietary lipids in endoge-
include the generation of homogeneous dispersions nous micellar species in the GI tract. In the simplest
with a high degree of drug-carrier miscibility, identi- sense, coadministration of drugs with lipids in lipid-
fication of systems with high glass transition temper- based formulations (LBF) seeks to harness the ad-
atures relative to storage temperatures, and the vantages of endogenous lipid processing pathways to
reduction of structural relaxation below the glass support drug absorption.
transition temperature. The stability of a solid disper- It has been recognized for many years that lipids in
sion is also dictated, at least in part, by the intrinsic food can assist the absorption of drugs with low water
crystallization tendencies of the drug—properties that solubility. Perhaps the best known example, and cer-
are difficult to predict a priori. tainly one of the earliest examples that benefit from
The choice of excipients that are used to formulate coadministration with food, is griseofulvin (Crounse,
solid dispersions remains empirical in many cases, and 1961). Others—including danazol (Charman et al.,
although experimentally determined indicators of 1993; Sunesen et al., 2005), halofantrine (Humberstone
utility, such as the long-held suggestion that storage et al., 1996), atovaquone, and troglitazone (Nicolaides
at 50°C below Tg results in acceptable stability, have et al., 2001; Schmidt and Dalhoff, 2002) have sub-
increased confidence in the progression of high-energy sequently been reported. Clinical reliance on coadmin-
formulations, it is clear that these provide an incom- istration with food is, however, fraught with difficulty
plete understanding of behavior in the glass state. because of the variability associated with differences in
Indeed, crystallization has been shown at temper- patterns of food ingestion (and food components) as
atures far below Tg, and it is increasingly apparent a function of a variety of factors, including time, age,
that crystallization behavior is not well correlated with culture, and health status. In contrast, coadministra-
global measures of structural relaxation and that tion of poorly water-soluble drugs with formulated
inferring changes in mobility below Tg from studies lipids provides a route to the advantages of lipid coad-
conducted above Tg may not be ideal. ministration, without the variability that results from
Instead, recent studies suggest that local molecular the use of food as a lipid source. The use of lipids in
motions such as Johari-Goldstein relaxations act as food as a mechanism of absorption enhancement is also
a precursor to changes in global mobility and insta- impractical in many preclinical models where animals
bility and might constitute a more accurate indicator of will not eat “on command.”
446 Williams et al.

The potential for LBF to enhance the absorption of Here we provide an overview of the mechanisms by
poorly soluble drugs has been recognized for more which LBF promote the absorption of poorly water-
than 40 years. Groves and de Galindez (1976) de- soluble drugs, describe the range of potential formula-
scribed LBF with the potential to self-emulsify within tions that fall under the broad umbrella of “LBF”,
the gastric environment in the 1970s, effectively discuss the design of LBF and the methods of for-
paving the way for the development of the encapsu- mulation assessment, and provide a number of recent
lated, self-emulsifying formulations that are most studies that exemplify the potential utility of this
commonly found in current use. Indeed, most contem- approach.
porary LBF are not simply solutions or suspensions of
drug in lipids but rather comprise combinations of A. Mechanisms of Bioavailability Enhancement by
glyceride lipids, lipophilic or hydrophilic surfactants, Lipid-Based Formulations
and cosolvents. These materials are usually filled into LBF are most commonly used to promote the oral
soft or hard gelatin capsules and may be liquid, bioavailability of poorly water-soluble drugs via
semisolid, or solid at room temperature. enhancements in drug solubilization and dissolution
The utility of LBF and the possible advantages of rate. In addition, LBF can stimulate intestinal lym-
self-emulsifying formulations are perhaps best exem- phatic drug transport (thereby providing increases in
plified by examination of the early formulations of the bioavailability via a reduction in first-pass metabo-
poorly water-soluble immunomodulator cyclosporine. lism), and emerging evidence suggests a potential role
The first of these products, Sandimmune, was mar- for components of LBF (primarily surfactants) in the
keted in 1981. Sandimmune comprised a solution of inhibition of intestinal efflux and metabolism. With the
cyclosporine in olive oil, Labrafil M 2125 CS, and exception of direct effects on transporter and enzyme
ethanol filled into soft gelatin capsules. Capsule functionality, the mechanisms by which LBF achieve
rupture in the stomach resulted in the formation of a these ends are intrinsically linked to their intercala-
polydisperse oil-in-water emulsion. The oral bioavail- tion into, or stimulation of, endogenous lipid digestion
ability of cyclosporine from Sandimmune was low but and absorption pathways. These are described in the
was able to provide for effective exposure after oral following section.
administration and was marketed for many years. In 1. Stimulation of Intestinal Lipid Absorption and
1994, Sandimmune was reformulated as Sandimmune Transport Pathways. Several aspects of intestinal
Neoral, comprising corn oil glycerides, polyoxyl 40 lipid absorption have been reviewed, in many cases
hydrogenated castor oil (Cremophor RH 40), propylene with a focus on strategies to reduce the rising incidence
glycol, ethanol, and DL-a-tocopherol. Unlike the origi- of metabolic syndrome. The specific topics addressed
nal Sandimmune formulation, Neoral was designed include lipid digestion (Phan and Tso, 2001; Mu and
such that it emulsified spontaneously on contact with Hoy, 2004; Lindquist and Hernell, 2010), nanostrucutres
the GI fluids to form a microemulsion with a mean formed during lipid digestion (Boyd, 2012), membrane
particle size smaller than 100 nm. The Neoral micro- transport (Niot et al., 2009), intracellular trafficking of
emulsion formulation resulted in enhanced and more lipids (Mansbach and Siddiqi, 2010), lipoprotein as-
consistent exposure and a reduction in the variability sembly (Iqbal and Hussain, 2009; Mansbach and
associated with coadministered food (i.e., a reduced Siddiqi, 2010; Warnakula et al., 2011), and lymphatic
food effect) (Kovarik et al., 1994; Mueller et al., 1994a, transport (Dixon, 2010). The interested reader is di-
b,c). The clinical and commercial success of Neoral rected to the listed references for more detail; however,
stimulated a subsequent increase in interest in the use the aspects most critical to drug absorption are sum-
of LBF and ultimately led to the development of LBF marized as follows.
for a range of drugs (Strickley, 2004, 2007). The success Dietary lipid is largely composed of triacylglyceride
of Sandimmune Neoral also resulted in significant or triglyceride (TG). Other dietary lipids include
interest in LBF that spontaneously form emulsions phospholipids (PL), sterols, glycolipids, and vitamins
with particle sizes in the nanometer size range on A, D, E, and K. Although TG is consumed in quantities
dispersion. Interestingly, although many of these of up to 100 g daily as part of a Western diet, it is
subsequently were to provide for excellent exposure, digested and absorbed with remarkable (95%) effi-
a direct link to the particle size of the initial dispersion ciency. The process of digestion and absorption is
remains elusive. Indeed, rapid processing of the initiated almost immediately on ingestion, as TG is
formulation in vivo (digestion, interaction with biliary partially hydrolyzed by lingual lipase, which is re-
components) likely leads to significant change in leased from lingual serous glands on the tongue, and
particle size. More recent data suggest that a better subsequently by gastric lipase in the stomach (Mu and
indicator of in vivo utility is the ability to maintain Hoy, 2004). Approximately 10–30% of dietary TG is
drug solubilization as the formulation is dispersed and hydrolyzed by gastric and lingual lipase to form
digested in the GI tract, a property also displayed by diglyceride (DG) and free fatty acid (FA) (Hamosh
the Neoral formulation. and Scow, 1973). The production of amphiphilic lipid
Strategies for Low Drug Solubility 447

digestion products, such as DG and FA, and the me- the surface of TG droplets as vesicular structures (or
chanical mixing that occurs in the stomach (Nordskog other dispersed liquid crystalline structures) and are
et al., 2001) facilitates the formation of a coarse lipid ultimately solubilized into mixed micellar structures
emulsion (droplet size ,50 mm) that increases the sur- comprising biliary derived PL, BS, and cholesterol. The
face area of exposure for further digestion. formation of a mixed micellar phase during lipid diges-
Entry of this coarse emulsion (chyme) into the tion was first described by Hofmann and Borgström
duodenum via the pylorus subsequently triggers (Hofmann and Borgström, 1964). Subsequent studies
a series of events that lead to the chemical digestion have identified myriad liquid crystalline species that
and physical solubilization of dietary lipids and exist in equilibrium with intestinal mixed micelles at
ultimately to highly efficient lipid absorption, mostly various stages during digestion of food or formulation-
from the upper small intestine (Nordskog et al., 2001; derived lipids (Patton and Carey, 1979; Patton et al.,
Mu and Hoy, 2004). Thus, the presence of TG digestion 1984; Rigler et al., 1986; Hernell et al., 1990; Staggers
products in the duodenum stimulates cholecystokinin et al., 1990; Fatouros et al., 2007a,b; Warren et al.,
release, gallbladder contraction, and bile and pancre- 2011). Progression through to mixed micelles, with
atic enzyme secretion into the duodenum (Nordskog small particle sizes (;3.5 nm) (Cohen and Carey, 1990)
et al., 2001; Mu and Hoy, 2004). Undigested lipid in the and high surface area-to-mass ratios is thought to
lower small intestine can also activate an “ileal brake” facilitate effective diffusional transport to the absorp-
that reduces small intestine motility and thus tive membrane.
increases the time available for further digestion and Integration of digested lipids into colloidal structures
absorption. In the duodenum, pancreatic lipase binds is required to maintain solubilization and at the same
to the surface of chyme and continues the digestion of time to provide structures that are small enough to
TG and DG to monoglyceride (MG) and free FA (Lowe, diffuse rapidly across the unstirred water layer (UWL)
1997). In the presence of bile salts, co-lipase is required present on the surface of absorptive enterocytes
to promote anchoring of lipase to the lipid droplet (Wilson et al., 1971; Nordskog et al., 2001). The UWL
interface (Lowe, 1997). Other dietary lipids are also is formed by water molecules trapped in a complex
hydrolyzed in the duodenum. For example, phosphati- glycoprotein network consisting of mucus and glyco-
dylcholine is digested by pancreatic phospholipase calyx and has a lower (acidic) pH compared with the
(PLA-2) to yield lysophosphatidylcholine (LPC) and bulk fluid in the intestine (Shiau et al., 1985; Orsenigo
free FA (Borgstrom et al., 1957; van den Bosch et al., et al., 1990). Historically, FA absorption was believed
1965; Arnesjo and Grubb, 1971) and cholesterol ester, to occur via passage across the UWL within mixed
which accounts for 10–15% of dietary cholesterol (with micelles and subsequent diffusion across the absorp-
the remainder present as free cholesterol) (Nordskog tive membrane after release from the mixed micelles in
et al., 2001), is hydrolyzed by cholesterol esterase. a process stimulated by the low pH microenvironment
Cholesterol esterase also has additional activity to- in the UWL (Shiau, 1981). In a seminal paper by Chow
ward TG and PLs (Lombardo et al., 1980). and Hollander, however, long-chain FA uptake across
Solubilization of lipids occurs simultaneously with the absorptive membrane was shown to occur via both
chemical digestion. Bile components—including bile passive and carrier-mediated transport, depending on
salts, PL, and cholesterol—are released into the the FA concentration (Chow and Hollander, 1979).
duodenum in the form of micelles after gallbladder Subsequently, several lipid transporters have been
contraction. Bile components facilitate further emulsi- identified (Niot et al., 2009) that interact specifically
fication of chyme, leading to a reduction in lipid droplet with free lipid (e.g., FA, cholesterol, PL) (Endemann
size and an increase in the surface area available for et al., 1993; Febbraio et al., 2001; Krieger, 2001;
lipase-mediated hydrolysis. Although the products of Thuahnai et al., 2001; Rajaraman et al., 2005). In some
lipid digestion are somewhat more water-soluble than cases, endocytic uptake of larger lipid complexes has
their glyceride precursors, they nonetheless have poor also been suggested (Pohl et al., 2002, 2004; Ring et al.,
bulk water solubility and as such accumulate on the 2002). However, unequivocal evidence of a role for
surface of digesting lipid droplets. As they are hydrated endocytosis in colloidal lipid uptake in the intestine
at the oil-water interface, these polar oils swell to form remains elusive.
liquid crystal structures ranging from lamellar to cubic Putative brush-border lipid membrane transporters
and hexagonal phase structures. The phase behavior is include cluster of differentiation 36 (CD36), plasma
complex and a function of the types and quantities of membrane fatty-acid binding protein (FABPpm), fatty-
lipids and lipid digestion products present and the acid transport protein 4 (FATP4), Niemann-Pick C1
concentration of bile salts, phospholipids, and other Like 1 (NPC1L1), and Scavenger receptor class B
dietary/formulation components (Patton and Carey, member 1 (SR-B1). A cholesterol antitransporter or
1979; Hernell et al., 1990; Staggers et al., 1990; efflux protein ABCG5/8 is also present (Berk and
Kossena et al., 2003; Fatouros et al., 2007a,b). In Stump, 1999; Niot et al., 2009). The quantitative
general, lipolytic products (FA, MG, DG) detach from importance of these transporters to plasma membrane
448 Williams et al.

uptake of lipids is controversial; however, there is consumed (Black, 2007; Mansbach and Gorelick, 2007;
strong evidence that each affects the metabolic fate of Iqbal and Hussain, 2009; Mansbach and Siddiqi, 2010).
lipids inside enterocytes (Berk and Stump, 1999; Niot Very-low-density lipoproteins (VLDLs) are preferen-
et al., 2009). Intracellular solubilization and transport tially produced in the fasted state, whereas both VLDL
of lipid digestion products also appears to be assisted and larger, less dense chylomicrons are formed after
by interaction with a number of families of lipid- lipid ingestion.
binding proteins [e.g., fatty-acid binding proteins Immature lipoproteins are subsequently trafficked
(FABP), acyl carrier proteins (ACP), sterol carrier pro- to the Golgi within prechylomicron transport vesicles
teins (SCP), cellular retinol binding proteins (CRBP)] (PCTVs) (Kumar and Mansbach, 1999; Mansbach and
(Agellon et al., 2002; Niot and Besnard, 2004; Niot Siddiqi, 2010), matured (Niot et al., 2009) and trans-
et al., 2009). Interaction with intracellular lipid- ferred to the basolateral membrane within vesicles
binding proteins has been suggested to facilitate containing multiple lipoproteins (Niot et al., 2009;
cytosolic transport and also to direct delivery to Mansbach and Siddiqi, 2010). These vesicles fuse with
intracellular organelles, including the nucleus (to allow the basolateral membrane, releasing lipoproteins into
genetic moderation of lipid absorption via changes to the intercellular space between enterocytes (Kindel
the expression of enzymes and proteins involved in et al., 2010; Kohan et al., 2010). After passing through
lipid absorption) and the endoplasmic reticulum, the basement membrane, lipoproteins are taken up
where re-esterification to TG and assembly into intes- into lacteals and the lymphatic system rather than in-
tinal lipoproteins occurs (Niot et al., 2009; Mansbach testinal microvessels. The specific mechanism of access
and Siddiqi, 2010). of lipoproteins into lacteals is the subject of some
The absorption and intracellular trafficking of lipids debate and is excellently reviewed by Dixon (Dixon,
are thus controlled by a highly specific network of 2010). Nonetheless, numerous studies have shown
transport proteins and metabolic enzymes. Simplisti- chylomicron entry into lacteals either via paracellular
cally, FA and MG with medium-length acyl chains, transport through open intercellular gaps (Palay and
which are more polar than their long-chain equiva- Karlin, 1959; Casley-Smith, 1962; Papp et al., 1962;
lents, diffuse across the enterocyte, enter the underly- Ladman et al., 1963; Rubin, 1966; Tytgat et al., 1971)
ing lamina propria, and access the systemic circulation or via transcellular transport within vesicles (Palay
via the intestinal vascular microvessels and portal vein and Karlin, 1959; Papp et al., 1962; Dobbins and
(Mansbach et al., 1991). In contrast, long-chain FA Rollins, 1970; Dobbins, 1971; Dixon et al., 2009). The
and MG are in most cases trafficked to the endoplas- mesenteric lymph vessels that drain the small in-
mic reticulum (ER), where they are resynthesized testine, converge to form larger lymphatic vessels, and
to TG via either the 2-monoglyceride pathway (which ultimately flow into the major lymphatic (the thoracic
predominates in the fed state) (Niot et al., 2009; lymph duct), which drains directly into the systemic
Mansbach and Siddiqi, 2010) or glycerol-3-phosphate blood circulation via a union with the vasculature at
pathway (which predominates in the fasted state) the junction of the left internal jugular and left sub-
(Johnston and Rao, 1965; Tso et al., 1995; Lehner clavian veins. This provides an access route for
and Kuksis, 1996). Synthesized TG then either enters lipoproteins to the systemic circulation and also pro-
a cytosolic pool (from where it may diffuse into vides a potential alternate transport pathway for
underlying intestinal microvessels or be hydrolyzed to poorly water-soluble, lipophilic drugs to the systemic
MG and FA and re-enter TG synthetic pathways) circulation. The latter provides a mechanism of drug
(Mansbach et al., 1991) or is transferred across the ER transport from the small intestine that bypasses the
membrane by microsomal TG transfer protein (MTP) liver and therefore has the potential to avoid hepatic
(Jamil et al., 1995; Hussain et al., 2003, 2008) and/or first-pass metabolism.
carnitine palmitoyltransferase (Washington et al., In addition to the traditional mechanisms of lipid
2003). MTP also facilitates lipidation of apolipoprotein transport to the systemic circulation described above
B48 (apoB48) with PL and cholesterol from the rough (i.e., diffusion directly into the mesenteric vasculature,
ER membrane to release a “primordial lipoprotein” or resynthesis, lipoprotein assembly, and uptake into
into the rough ER lumen (Wu et al., 1996; Hussain mesenteric lacteals), an alternative pathway has been
et al., 2008). Fusion of the primordial lipoprotein with relatively recently described, where cholesterol and
the TG droplet at the junction of the smooth ER and vitamin D are secreted from the enterocyte within
rough ER leads to formation of an immature lipopro- apoB free/ApoAI containing lipoproteins that are
tein or prechylomicron via a process termed core essentially identical to high-density lipoprotein (HDL)
expansion (Black, 2007; Mansbach and Gorelick, (Iqbal et al., 2003; Iqbal and Hussain, 2005). The
2007; Iqbal and Hussain, 2009; Mansbach and Siddiqi, subsequent fate of HDL released from the intestine,
2010). The ultimate size of the lipoproteins formed is however, is also controversial (Iqbal and Hussain,
dictated by the size of the core lipid droplet, which in 2009), with some authors proposing direct uptake into
turn depends on the quantity and type of lipid the mesenteric lymphatics (Bearnot et al., 1982;
Strategies for Low Drug Solubility 449

Forester et al., 1983; Magun et al., 1985), whereas colloidal species, including emulsion droplets,
others suggest that HDL in lymph is derived from vesicles, and mixed micelles that have high
plasma (Quintao et al., 1979; Sipahi et al., 1989; solubilization capacities for poorly water-soluble
Oliveira et al., 1993; Brunham et al., 2006). molecules
The absorption of poorly water-soluble dietary lipids • Promotion of mass transport across the unstirred
and subsequent transport to the systemic circulation is water layer via transport within intestinal mixed
therefore supported via a number of processes. These micelles
include the following: • Absorption across the apical (absorptive) mem-
brane of the enterocyte via passive and active
• Stimulation of bile secretion and an increase in transport mechanisms
intestinal concentrations of biliary-derived • Transport across the enterocyte via interaction
solubilizers with cytosolic lipid-binding proteins
• The combination of biliary secretions with lipid • Assembly into intestinal lipoproteins within the
digestion products in the intestinal lumen to form enterocyte, uptake into the intestinal lymph, and

Fig. 63. Schematic of the mechanisms by which poorly water-soluble drugs are absorbed from lipid-based formulations (LBF), including illustration of
the mechanisms by which LBF promote the absorption or bioavailability of poorly water-soluble drugs. (1) Avoidance of the need for traditional drug
dissolution by presenting the drug as a monomolecular dispersion (solution) in the LBF. (2) Formation of a dispersed emulsion within the stomach in
which the drug remains solubilized after capsule rupture. (3) Stimulation of bile secretion, bile salt, and phospholipid micelle release into the intestinal
lumen and thus an increase in intestinal concentrations of biliary-derived solubilizers. (4) Combination of biliary secretions with lipid digestion
products in the small intestine lumen to form colloidal species—including emulsion droplets, vesicles, and mixed micelles—that have high
solubilization capacities for poorly water-soluble drugs. Note that the release of pancreatic lipase and co-lipase from the gallbladder facilitates
digestion of formulation lipids and thus the formation of more polar digestion products that assist in vesicle and mixed micelle formation. (5) Promotion
of mass transport across the unstirred water layer (UWL) via transport within intestinal mixed micelles. (6) Enhanced drug permeability across the
enterocyte apical membrane where formulation components may alter passive permeability or reduce efflux. (7) Transport across the enterocyte via
interaction with cytosolic lipid-binding proteins. (8) Highly lipophilic drugs (typically log D . 5, long-chain TG solubility . 50 mg/g) have the potential
for association with developing lipoproteins in the enterocyte and uptake into the intestinal lymph rather than the mesenteric blood capillaries, and
therefore transport to the systemic circulation via a route that avoids passage through the liver and first-pass hepatic metabolism. (9) Protection from
enterocyte-based drug metabolism, either directly via excipient effects on enzyme activity or indirectly where lipoprotein association within the
enterocyte occurs.
450 Williams et al.

transport to the systemic circulation in a transport results in drug precipitation in the amorphous state
route that avoids the liver (Sassene et al., 2010; Thomas et al., 2012), both
situations that are expected to enhance dissolution rate.
The coadministration of lipophilic drugs with lipids, Lipid suspension formulations have also been used
therefore, has the potential to recruit a number of as LBF (Bloedow and Hayton, 1976; Kaukonen et al.,
endogenous lipid processing pathways to support drug 2004a; Dahan and Hoffman, 2007; Jantratid et al.,
absorption. These can be broadly grouped into mech- 2008; Larsen et al., 2008a; Sachs-Barrable et al., 2008),
anisms that support solubilization, those that facilitate especially where drug loads are sufficiently high to
absorption and permeability across the enterocyte, and preclude the generation of a solution formulation. In
those that stimulate intestinal lymphatic transport. this case, the benefits of LBF are limited to their
These mechanisms are depicted in Fig. 63. In addition ability to raise the solubilization capacity of the GI
to lipids, LBF also often contain other excipients (e.g., fluids and must rely on dissolution for the drug dose to
surfactants and cosolvents) that may further assist in be solubilized in situ. In this case, exposure profiles can
drug solubilization, absorption, or lymphatic transport. be more variable than that achieved with lipid solution
The mechanisms by which LBF support absorption are formulations. Processing and capsule filling complex-
discussed below realizing that benefits accrue as ities further confound the development of suspension
a function of both stimulation of endogenous lipid formulations.
processing pathways and the additional benefits that Solubilization of a lipophilic, poorly water-soluble
exogenous, formulation-derived materials provide. drug in the GI content is therefore a function of the
2. Enhanced Drug Dissolution and Solubilization in following:
the Intestinal Lumen. LBF typically comprise lipids • Avoidance of drug precipitation as LBF are
or combinations of lipids, surfactants, and cosolvents initially dispersed (usually on capsule rupture in
that are chosen such that the drug dose can be the stomach)
completely dissolved in the combination of excipients. • Stimulation of secretion of endogenous solubiliz-
In most cases, this lipidic solution is filled into hard or ing lipids (bile salt, phospholipid, cholesterol) in
soft gelatin capsules to allow convenient administra- bile
tion. On capsule rupture, LBF therefore present drug • Supply of exogenous (formulation-derived) lipids,
to the GI tract in solution, albeit in nonaqueous lipid digestion products, surfactants, and
solution, providing an immediate benefit compared cosolvents
with traditional solid dose forms since it avoids the • Generation of mixed colloidal species from endog-
requirement for wetting and dissolution. Instead, LBF enous and exogenous solubilizers with the solubi-
substitute dissolution for the requirement for contin- lization capacity to prevent drug precipitation
ued solubilization as the formulation is processed in during GI transit
the GI tract and the need for transfer or partition into
colloidal droplets that are sufficiently small to allow These processes are depicted in Fig. 63.
diffusion across the UWL and approach to the The likelihood of drug precipitation during formula-
absorptive membrane. At the absorptive membrane, tion dispersion is described in more detail in Sections
current understanding suggests that drug absorption XI.A.2 and XI.B.3. Factors affecting crystallization in
occurs via the fraction of drug that is present in free oil-in-water emulsions have been reviewed in detail
solution and in equilibrium with the solubilized (McClements, 2012). Simplistically, drug precipitation
reservoir. Drug absorption depletes the free concentra- is largely dictated by the hydrophilicity of the compo-
tion in solution, and this is rapidly replenished by drug nents present in the formulation. Thus, simple solu-
partition out of the solubilized reservoir to maintain tions of drug in lipids (the most lipophilic of LBF)
the solubilization equilibrium. In essence, the tradi- carry little threat of drug precipitation as the formu-
tional solid-liquid dissolution step that occurs with lation is dispersed (Porter and Charman, 2001a;
solid dose forms (and limited by the solid-state Abdalla and Mäder, 2009; Mohsin et al., 2009; Day
properties of the crystal lattice) is replaced by a rapid et al., 2010). In contrast, incorporation of larger quan-
liquid-liquid partitioning step between the solubilized tities of water-miscible surfactants and cosolvents
reservoir and drug in free solution. Continued solubili- increases the risk of precipation as the solubilizing
zation is important, as drug precipitation recreates solid power of these components is lost on dilution. In
drug and results in reversion of the drug absorption general, less lipophilic drugs appear to more favorably
process to one that reflects the situation on administra- partition out of digesting lipidic formulations and into
tion of a solid dose form or suspension. As a caveat to the colloidal lipid phases formed by the incorporation
this general assumption, however, some benefits may be of digestion products into endogenous micellar species
evident if precipitation leads to the generation of small (Kaukonen et al., 2004a; Day et al., 2010). In contrast,
drug particles or where phase separation from a LBF more lipophilic drugs accumulate in any undigested oil
Strategies for Low Drug Solubility 451

that is present, limiting transfer into micellar compo- ionized MC FAs, such as capric and caprylic acid, are
nents until all remaining lipid is digested (Sek et al., significantly more polar and much less capable of
2002; Kaukonen et al., 2004a; Day et al., 2010). As swelling bile salt micelles (and therefore of increasing
expected, drugs with higher affinities for bile salt and the solubilization capacity) than the equivalent glyc-
phospholipid mixed micelles equilibrate into the erides. In contrast, the digestion products of LC glyc-
colloidal lipid phase more effectively (Christensen erides, for example, oleic acid, are much more capable
et al., 2004). of retaining the capacity to support drug solubilization
Solubilization is further influenced by the mass and on digestion. The utility of LC lipids has been con-
type of lipid administered in the formulation, and, in firmed in several in vivo studies, where more effective
general, administration of a higher lipid mass provides support of drug absorption is evident compared with
greater support for drug solubilization. This reflects MC lipids (Palin and Wilson, 1984; Holmberg et al.,
the ability of larger quantities of lipid to maximize the 1990; Behrens et al., 1996; Christensen et al., 2004;
release of bile salt (Kossena et al., 2007) and also the Porter et al., 2004b; Han et al., 2009). This is not
ability of larger quantities of lipid digestion products to always the case, however, and others have described
more effectively swell endogenous bile salt micelles, similar absorption from MC and LC systems (Grove
thereby increasing solubilization capacity. Effective et al., 2005, 2006) or benefits associated with the use of
differentiation of the ability of different lipids to most MC systems (Gallo-Torres et al., 1978; Myers and
effectively promote drug solubilization in the GI tract Stella, 1992; Dahan and Hoffman, 2006, 2007; Ahmed
remains a goal of many research programs, including et al., 2012).
ours; however, it is apparent that the inclusion of, for The inclusion of surfactants and cosolvents in LBF
example, glycerides based on medium-chain (MC) or has a profound impact on drug solubilization, both on
long-chain (LC) FA leads to quite different colloidal formulation dispersion and on digestion. As described,
behavior and, therefore, differences in drug solubiliza- increased quantities of hydrophilic materials in LBF
tion. For example, Kossena et al. (2003, 2004) ex- promote the formation of emulsions with smaller
amined phase behavior and drug solubilization on particle size (often in the nanometer size range) but
dilution of MC and LC FA and monoglyceride (MG) also increase the risk of drug precipitation during
(reflecting common digestion products) in simulated dispersion owing to a loss of solubilization capacity
endogenous intestinal fluid (SEIF: 4 mM BS, 1 mM PL, Pouton, 2006; Cuiné et al., 2007; Mohsin et al., 2009).
and 0.25 mM cholesterol). At high concentrations of This is particularly true for cosolvents, the solubiliza-
lipid, reflective of conditions likely to exist on the tion capacity of which decreases exponentially on
surface of a digesting oil droplet, liquid crystalline dilution [see Section VI.E and Pouton, 2006; Pole,
phases dominated with a lamellar phase evident in MC 2008]. In contrast, LBF containing less water-soluble
FA and MG systems and a more viscous cubic phase surfactants are less able to generate emulsions with
formed in the long-chain FA and MG systems. Further very small particle size but are more resistant to drug
dilution resulted in the production of large vesicles and precipitation on dispersion.
the coexistence of mixed micelles and unilamellar One aspect of the performance of formulations
vesicles at concentrations likely to reflect the environ- containing surfactants after oral administration that
ment adjacent to the absorptive surface of the enterocyte. is often overlooked is the potential for surfactant dig-
The MC systems were vesicle rich and had higher estion in vivo. Many of the surfactants used in phar-
solubilization capacities at high FA and MG concen- maceutical preparations are FA esters of ethoxylated
trations compared with the LC systems; however, the hydrophilic head groups (see Section VII) and as
solubilization capacity of the MC system was rapidly such have the potential for enzymatic cleavage of the
lost on dilution. In contrast, the LC systems retained ester bond. Surfactants such as polysorbate 80,
solubilization capacity through greater levels of di- Cremophor EL, and Cremophor RH 40 (Cuiné et al.,
lution, coincident with the persistence of vesicular 2008; Christiansen et al., 2010) (although Cremophor
species at greater dilutions. Therefore, LC lipids ap- RH40 may be less effectively hydrolyzed than Cremo-
pear to support more robustly drug solubilization at phor EL (Cuiné et al., 2008)), Labrasol, and Gelucire
high dilution. (Cuiné et al., 2008; Fernandez et al., 2008, 2009) are
It is also apparent that the properties of formulation- digestible, whereas TPGS and sucrose laurate appear
derived lipids change significantly on digestion. In- to be stable in the presence of pancreatic enzymes
deed, differential solubilization properties may become (Christiansen et al., 2010). Whether digestion affects
apparent only on digestion. For example, drug solubil- the properties of surfactants sufficiently to alter the
ity in MC lipids is usually higher (on a milligram per solubilization capacity of the system, however, is less
gram basis) than in LC lipids, providing advantages in well understood. Digestion of surfactants leads to a loss
drug solubility in the formulation and therefore the of solvent capacity (Cuiné et al., 2008; Fernandez et al.,
potential for the administration of higher drug doses in 2008,2009) during in vitro lipolysis, and this finding
the same mass of formulation. On digestion, however, has also been correlated with reductions in absorption
452 Williams et al.

in vivo in at least one study (Cuiné et al., 2008). The use precipitation. It is currently not well defined across
of nondigestible surfactants has therefore been sug- different formulation classes and drug types. Nonethe-
gested to provide an avenue to the generation of more less, supersaturation is likely to occur when LBF lose
robust LBF (Cuiné et al., 2008). Surfactants can also solubilization capacity during GI processing [see
inhibit the activity of pancreatic lipase, although it is Section XI.B.3] such as during dilution and dispersion
uncertain how relevant this is to the in vivo processing of formulations with higher concentrations of surfac-
of LBF where enzyme levels are high and probably in tant or cosolvent (Mohsin et al., 2009; Do et al., 2011)
excess (MacGregor et al., 1997; Sek et al., 2006; or during digestion of glyceride-containing formula-
Christiansen et al., 2010; Li and McClements, 2011; tions (Kaukonen et al., 2004b; Anby et al., 2012;
Wulff-Pérez et al., 2012). Williams et al., 2012b). Supersaturation might also
The utility of LBF is therefore, at least in part, be expected to occur more readily in MC LBF (which
a function of the ability to maintain drug in a solubi- lose solubilization capacity more quickly during dis-
lized state during formulation dispersion, digestion, persion and digestion) compared with LC systems. As
and interaction of exogenous lipids and surfactants such, supersaturation ratios may be particularly high
with endogenous solubilizing species within the GI for highly lipophilic drugs after digestion of MC
fluid. Interestingly, however, recent studies have systems (Kaukonen et al., 2004b; Porter et al., 2004a;
shown that the generation of colloidal structures with Williams et al., 2012b). The challenge in this case is the
high equilibrium solubilities may not be critical to stabilization of supersaturation and prevention of
absorption promotion. For example, where drug re- precipitation.
lease from the LBF is slowed (e.g., via the formation of Strategies to assist in the stabilization of supersatu-
cubic liquid crystalline phases), absorption may be ration are therefore widely sought. In general, the same
enhanced by the release of constant low concentrations technologies as those used to stabilize supersaturation
of drug to the GI environment, thereby lowering the resulting from the dissolution of salts, high-energy
required solubilization capacity (Nguyen et al., 2010). crystal forms, or solid dispersions are typically used.
LBF dispersion and digestion may also lead to a period Thus, the addition of polymeric precipitation inhibitors
when drug remains solubilized but is present at has been used to good effect with LBF, although the av-
concentrations that are supersaturated compared with ailable database is relatively small (Gao and Morozowich,
equilibrium drug solubility in the colloidal species 2006; Brouwers et al., 2009; Warren et al., 2010; Anby
present. Where supersaturation is retained for a period et al., 2012; Li et al., 2012). Polymer-containing SEDDS
that is sufficient to support absorption, bioavailability systems have been referred to as “supersaturable” SEDDS
may be enhanced despite only moderate changes to (Erlich et al., 1999; Gao et al., 2004; Gao and Morozowich,
equilibrium solubility (Porter et al., 2011; Anby et al., 2006; Gao et al., 2009, 2011a). A range of polymers
2012; Li et al., 2012; Thomas et al., 2012). The have been explored for their effect on supersaturation
generation of supersaturation may also provide addi- stabilization; however, to this point, cellulose-based
tional benefits with respect to absorption by increasing polymers appear to be most effective (Warren et al.,
drug thermodynamic activity. Supersaturation as a 2010). Polymers likely prevent precipitation by inter-
potential driving force for absorption was reviewed fering with crystal nucleation or growth (Gao and
(Brouwers et al., 2009; Augustijns and Brewster, 2012) Morozowich, 2006; Brouwers et al., 2009; Warren et al.,
and more specifically in the context of LBF by Warren 2010). Even in the event of precipitation, polymers may
et al. (2010) and Gao and Morozowich (2006). also retain some benefit by encouraging drug pre-
The length of time for which drug must remain cipitation in the amorphous rather than the crystalline
solubilized to promote drug absorption effectively is form, providing the potential for enhanced resolubili-
drug specific and reflects intestinal permeability. As zation (Gao et al., 2009).
such, solubilization may need to be maintained only 3. Enhanced Intestinal Permeability and Inhibition
briefly for highly permeable drugs but retained for of Intestinal Efflux and First-Pass Metabolism.
longer periods for less permeable compounds where The high lipophilicity and limited polarity of poorly
absorption is slower. The likely period of supersatura- water-soluble drugs dictate that most have intrinsi-
tion depends partly on the mass ratio of drug solubilized cally good passive membrane permeability (i.e., they
in the supersaturated condition to the equilibrium behave as typical class II BCS compounds). However,
solubility of the drug in the system (i.e., the supersat- this is not always the case, and several examples of
uration ratio) (Augustijns and Brewster, 2012). In low-solubility, low-permeability compounds are evident
general, the likelihood of crystal nucleation and pre- (BCS class IV). Poorly water-soluble, highly lipophilic
cipitation is enhanced at higher supersaturation ratios drugs are also commonly substrates for intestinal efflux
(Gao and Morozowich, 2006; Warren et al., 2010; Anby transporters such as P-glycoprotein (P-gp), multidrug-
et al., 2012; Augustijns and Brewster, 2012). The resistance proteins (MRP) and breast cancer resistance
optimum supersaturation ratio is, however, a balance protein (BCRP) and have an increased liability toward
between enhanced solubilization and the potential for first-pass metabolism. The challenges provided by low
Strategies for Low Drug Solubility 453

water solubility, therefore, often occur in tandem with Bilayer curvature strain occurs when a surfactant
issues of permeability, especially cellular efflux or inserts quickly into the outer monolayer but does not
metabolism. Under these circumstances, formulation translocate to the inner leaflet or flip-flops relatively
components that assist solubility and at the same time slowly, resulting in bilayer asymmetry and strain. Flip-
enhance permeability or reduce first-pass metabolism flop requires the polar headgroup of the surfactant to
are beneficial. traverse the hydrophobic core of the membrane and to
Many components present within LBF, including align with the polar headgroups of the inner rather
triglyceride digestion products (FA and MG), surfac- than the outer membrane leaflet. As such, surfactants
tants, cosolvents, and biliary components (bile salts, with large, charged, or very hydrophilic head groups
lysophospholipids, cholesterol) enhance passive para- are more likely to flip-flop slowly and to generate bi-
cellular and transcellular membrane diffusion (Aungst, layer curvature strain. If the bilayer is unable to
2000; Seelig and Gerebtzoff, 2006; Heerklotz, 2008; assume the spontaneous curvature required to accom-
Goole et al., 2010). For example, MC FAs (sodium modate asymmetry, strain occurs as a result of mole-
caprate) and glycerides (C8–C10) (Sekine et al., 1985; cular compression. Where the surfactant head group is
Unowsky et al., 1988; Yeh et al., 1994) are well known large and the hydrophobic surfactant tail is sufficiently
to enhance paracellular diffusion by opening tight small that it cannot completely fill the void between
junctions (Anderberg et al., 1992; Tomita et al., 1992; adjacent membrane lipids, monolayer curvature strain
Lohikangas et al., 1994; Lindmark et al., 1998) and also leads to a disordering of the hydrophobic chains in
may also permeabilize the membrane and enhance the membrane and a thinner, more flexible, and more
transcellular permeation (Tomita et al., 1992). Longer- permeable membrane with reduced lateral packing
chain FAs and MGs similarly enhance passive perme- density. Strong detergents are more able to enhance
ability, although fewer studies have examined the passive permeability and are considered those that
effects of LCs compared with MC FAs (Lundin et al., partition readily into membranes (generally those with
1997; Aungst, 2000). Surfactants and cosurfactants higher log P) and that have large head groups or small
also enhance paracellular diffusion by opening tight hydrophobic groups and therefore induce either mono-
junctions (Rama Prasad et al., 2003; Chang and layer or bilayer curvature strain. At higher surfactant
Shojaei, 2004; Rama Prasad et al., 2004; Sha et al., concentrations, bilayer or monolayer curvature strain
2005) or transcellular diffusion by permeabilizing or can result in mechanical failure of the membrane,
solubilizing the membrane (Rege et al., 2002; Prasad membrane leakage, and increased passage of solutes
et al., 2003; Rama Prasad et al., 2003). across the membrane.
The potential for a surfactant to increase trans- Surfactants also enhance passive drug diffusion via
cellular diffusion via an increase in membrane perme- membrane solubilization. Membrane solubilization re-
ability is related to structure and physicochemical sults when surfactants cause curvature strain to such
properties (Seelig and Gerebtzoff, 2006; reviewed in an extent that lipids are expelled from the membrane
Heerklotz, 2008) and is determined by differences in in the form of lipid-saturated surfactant micelles.
both uptake into the membrane and perturbation of Solubilization is more likely to occur with strong
membrane structure. Uptake into the outer membrane detergents that induce curvature strain at relatively
leaflet is usually rapid; however, flip-flop to the inner low concentrations and that are able to induce sig-
membrane leaflet may occur slowly for some surfac- nificant perturbations in the membrane without being
tants, in particular those with bulky, ionic, or highly squeezed out of the membrane by the induced
hydrophilic headgroups. Membrane uptake is governed curvature (typically surfactants that are anchored into
by the mole ratio partition coefficient, Kr, the membrane by more than a single alkyl chain). Bile
salts are particularly strong membrane solubilizers
nbs and are therefore very effective permeability enhancers
Kr 5 ð63Þ
nbL × Csaq in vitro (Scott Swenson and Curatolo, 1992; Aungst,
2000; Heerklotz, 2008). However, whether the concen-
where nbs and nbL are the mole numbers of surfactant and trations of lipids or bile salts that are likely to be pre-
lipid in the membrane bilayer, respectively, and Csaq is sent in the GI tract during the digestion of LBF are
aqueous surfactant concentration (Heerklotz, 2008). In sufficient to generate permeability enhancement in vivo
general, Kr is proportional to the log P of the surfactant is less clear. Assembly of bile salts and lipids into mixed
and approximates 1/CMC. Thus, more lipophilic surfac- micelles of bile salt, phospholipid, cholesterol, and
tants have lower CMCs and partition more readily into exogenous lipids also significantly reduces the thermo-
membranes. However, deviations from this relationship dynamic activity of the bile salts and reduces the
are possible, for example, where the surfactant is likelihood of permeability enhancement.
charged and forms electrostatic interactions with the Surfactants can therefore enhance passive diffusion
membrane or where the surfactant induces bilayer or across membranes via membrane permeabilization
monolayer curvature strain in the membrane. and solubilization. However, some degree of membrane
454 Williams et al.

TABLE 27
Summary of studies investigating the impact of lipid formulation components on drug efflux by major intestinal efflux transporters [P-glycoprotein
(P-gp), breast cancer resistance protein (BCRP), and multidrug resistance protein-2 (MRP-2)]
Solubilizer Transporter Concentration Effects

Brij 30 P-glycoprotein (Lo, 2003; 20–200 mM Inhibition of epirubicin or mitoxantrone


Yamagata et al., 2007a) transport in Caco-2 cells, P-gp expressing
cells and everted gut segments
BCRP (Yamagata et al., 50–100 mM Inhibition of mitoxantrone transport in
2007a) BCRP-expressing cells
Brij 35 P-glycoprotein (Yu et al., 30–120 mM Inhibition of bis(12)-hupyridone efflux
2011a) in Caco-2 cells
Cremophor EL P-glycoprotein (Batrakova In vitro ,2% (w/v) Inhibition of Rhodamine 123, mitoxantrone,
et al., 1998; Bogman et al., digoxin, verapamil or paclitaxel transport
2003; Cornaire et al., 2004; in Caco-2 cells, P388/MDR cells, P-gp
Hugger et al., 2002; Katneni overexpressing cells, rat intestinal tissue
et al., 2007; Martin-Facklam or everted gut segments
et al., 2002a; Mudra and In vivo 1 mg/kg rats, Altered pharmacokinetic profile of digoxin
Borchardt, 2010; Rege et al., 0–5000 mg in humans and celiprolol, and dose dependent increase
2002; Shono et al., 2004) (i.e., 0, 100, 1000, or in plasma Cmax and AUC of saquinavir,
5000 mg) after oral administration to humans
MRP-2 (Bogman et al., 2003) 0.0002–2% (w/v) No significant effect on methylfluorescein-
glutathione conjugate transport by MDCK
cells overexpressing MRP-2
BCRP (Yamagata et al., 2007a) 30–50 mM Inhibition of mitoxantrone transport in
BCRP-expressing cells
Cremophor RH40 P-glycoprotein (Tayrouz et al., In vitro 50–100 mM Inhibition of mitoxantrone transport in
2003; Yamagata et al., 2007a) P-gp overexpressing cells
In vivo 600 mg, 3 daily Altered pharmacokinetic profile of digoxin
after oral administration to humans
BCRP (Yamagata et al., 2007a) 50–100 mM Inhibition of mitoxantrone transport in
BCRP overexpressing cells
CRL-1605 copolymer P-glycoprotein (Banerjee et al., 132 mg/kg Altered pharmacokinetic profile of
2000; Jagannath et al., 1999) tobramycin and amikacin following
oral administration to mice
Gelucire 44/14 P-glycoprotein (Sachs-Barrable 0-0.02% (v/v), 1–15 mM Inhibition of mitoxantrone or rhodamine
et al., 2007; Yamagata et al., transport in Caco-2 cells and P-gp
2007a) overexpressing cells
BCRP (Yamagata et al., 2007a) 1–15 mM No difference in mitoxantrone uptake in
BCRP-expressing cells
Labrasol P-glycoprotein (Cornaire et al., 0.05–0.5% (w/v) Inhibition of rhodamine or digoxin and celiprolol
2004; Hu et al., 2002; Lin transport in vitro and in rat everted gut
et al., 2007b) segments
5 mg/kg into rat colon Altered pharmacokinetic profile of digoxin and
celiprolol in rat
bioavailability studies
Myrj 52 P-glycoprotein (Lo, 2003; 20–200 mM in vitro. Inhibition of epirubicin or mitoxantrone
Yamagata et al., 2007a) 200 mM in situ transport in vitro Caco-2 cells, P-gp
expressing cells and in situ everted
gut sacs
BCRP (Yamagata et al., 2007a) 100–200 mM No effect on mitoxantrone transport in
BCRP over-expressing cells
PEG 200, PEG 300, P-glycoprotein (Hugger et al., 0–20% Inhibition of doxorubicin, taxol and
PEG 400, and PEG 2002; Johnson et al., 2002; rhodamine 123 transport in Caco-2
20,000 Shen et al., 2006, 2008) cells and MDR1-MDCK cells
Inhibition of prednisolone,
methylprednisolone, quinidine and
digoxin transport in situ or excised
rat jejunum
Pluronic F68 P-glycoprotein (Batrakova et al., 0–2%, 0–7050 mM No significant effect on
1998; Bogman et al., 2003; bis(12)-hupyridone or rhodamine
Yu et al., 2011a) 123 transport in Caco-2 cells and
P-gp expressing cells, respectively
Inhibition of rhodamine 123 transport
in Caco-2 cells but potency less than
Pluronic P85 and L81
MRP-2 (Bogman et al., 2003) 0.0002–2% (w/v) No significant effect on
methylfluorescein-glutathione
conjugate transport in vitro MDCK
cells overexpressing MRP-2
Pluronic F108 P-glycoprotein (Miller et al., 100 mM Inhibition of rhodamine 123 transport
1999) in Kbv monolayers
MRP-2 (Miller et al., 1999) 100 mM Inhibition of fluorescein transport in
panc-1 cell monolayers
Pluronic L81 P-glycoprotein (Batrakova et al., 40–100 mM Inhibition of rhodamine 123 transport
1998; Miller et al., 1999) in Caco-2 cells and Kbv monolayers
MRP-2 (Miller et al., 1999) 100 mM Inhibition of fluorescein transport in
panc-1 cell monolayers
(continued )
Strategies for Low Drug Solubility 455
TABLE 27—Continued
Solubilizer Transporter Concentration Effects

Pluronic P85 P-glycoprotein (Batrakova et al., 0–220 mM Inhibition of mitoxantrone transport


1998; Miller et al., 1999; in P-gp expressing cells
Yamagata et al., 2007a) Inhibition of digoxin transport in situ
rat jejunum
MRP-2 (Miller et al., 1999) 100 mM Inhibition of fluorescein transport in
panc-1 cell monolayers.
BCRP (Yamagata et al., 2007a,b) In vitro 1–20 mM Inhibition of mitoxantrone transport
in BCRP-expressing cells
In vivo 100–500 mg/kg Dose dependent increase in topotecan
AUC in wild-type mice, maximum
twofold increase and no effect on
IV clearance. Less effective in
increasing AUC after administration
to BCRP (2/2) mice when compared
with BCRP (1/1) mice (1.2-fold
increase). Similar effects seen in
everted gut sacs
Pluronic PE8100 P-glycoprotein (Bogman 0.0002–2% (w/v) Inhibition of rhodamine 123 transport at
and Pluronic et al., 2003) in P388/MDR cells
PE6100
MRP-2 (Bogman et al., 2003) 0.0002–2% (w/v) No significant effect on methylfluorescein-
glutathione conjugate transport by MRP-2
in MDCK cells overexpressing MRP-2
Poloxamer 188 P-glycoprotein (Bogman 0.8% (w/v) solution No effect on talinolol transport in Caco-2
et al., 2005) cells or on absorption in vivo in humans
Propylene glycol P-glycoprotein (Yamagata 1–15 mM No significant effect on mitoxantrone
et al., 2007a) transport in P-gp expressing cells
BCRP (Yamagata et al., 1–15 mM No significant effect on mitoxantrone
2007a) transport in BCRP-expressing cells
Sodium lauryl sulfate P-glycoprotein (Bogman et al., 2003; 0.0002–2% (w/v) No significant effect on rhodamine 123
Lo, 2003; Shono et al., 2004) transport
20–200 mM Inhibition of epirubicin transport in
Caco-2 cells in a dose-dependent
manner. Effect less pronounced
than other excipients
MRP-2 (Bogman et al., 2003) 0.0002–2% (w/v) No significant effect on methylfluorescein-
glutathione conjugate transport in vitro
MDCK cells overexpressing MRP-2
Softigen 767 P-glycoprotein (Cornaire et al., 2004) In situ 0.05–0.5% (w/v) Inhibition of digoxin and celiprolol
In vivo 1–120 mg/kg transport in rat everted gut segments
Significant increase in digoxin plasma
AUC and modified plasma profile
of celiprolol in vivo in rats
Solutol HS P-glycoprotein (Bittner et al., 2002; In situ 0.05–0.5% (w/v). Inhibition of digoxin transport in rat
Cornaire et al., 2004) In vivo 10% (w/v) solution everted gut segments
2-fold increase in colchicine
bioavailability in rats
Span 20 P-glycoprotein (Yamagata et al., 2007a) 50–100 mM Inhibition of mitoxantrone transport
in P-gp expressing cells
BCRP (Yamagata et al., 2007a) 50–100 mM Inhibition of mitoxantrone transport in
BCRP-expressing cells
Span 80 P-glycoprotein (Lo, 2003) 20–200 mM Inhibition of epirubicin transport in
Caco-2 cells and in situ rat
everted gut sacs
TPGS P-glycoprotein (Bittner et al., 2002; In vitro 0–1.0 mM Inhibition of of Rhodamine 123, paclitaxel,
Bogman et al., 2003, 2005; mitoxantrone, and talinolol transport in
Brouwers et al., 2006; Chan et al., caco-2 cells and P388/MDR cells.
2004; Chang and Shojaei, 2004; Enhanced the cytotoxicity of P-gp
Chang et al., 1996; Dintaman and substrates (doxorubicin, vinblastine,
Silverman, 1999; Pan et al., 1996; paclitaxel, and colchicines) in
Rege et al., 2002; Sokol et al., 1991; NIH-MDR1-GA85 cells
Varma and Panchagnula, 2005; 0.0002–2% (w/v) Inhibition of paclitaxel transport but no
Yamagata et al., 2007a) effect on digoxin transport in situ
rat jejunal tissue
In vivo 50 mg/kg TPGS Increased amprenavir, colchicine,
cyclosporine, talinolol and paclitaxel
bioavailability in rats or humans
MRP-2 (Bogman et al., 2003) 0.0002–2% (w/v) No significant effect on
methylfluorescein-glutathione
conjugate transport in MDCK cells
overexpressing MRP-2
BCRP (Yamagata et al., 2007a) 50–100 mM No significant effect on mitoxantrone
transport in BCRP-expressing cells
Tween 20 P-glycoprotein (Batrakova et al., 20–250 mM Inhibition of mitoxantrone, epirubicin
1998; Cornaire et al., 2004; Lo, 2003; and rhodamine 123 transport in
Yamagata et al., 2007a) P-gp expressing cells and Caco-2 cells
(continued )
456 Williams et al.

TABLE 27—Continued
Solubilizer Transporter Concentration Effects

Inhibition of epirubicin and digoxin


transport in rat everted gut sacs
BCRP (Yamagata et al., 2007a,b) 100-250 mM Inhibition of mitoxantrone transport in
BCRP-expressing cells
Dose dependent increase in topotecan
AUC in wild type mice, maximum
1.8 fold increase and no effect on i.v.
clearance. Less effective in increasing
AUC after administration to BCRP (2/2) mice
(1.2-fold increase).
Similar effects seen in everted gut sacs
Tween 80 P-glycoprotein (Batrakova et al., In vitro 0–1 mM Inhibition of epirubicin, rhodamine 123,
1998; Bogman et al., 2003; bis(12)-hupyridone, mitoxantrone,
Cornaire et al., 2004; Hugger and taxol transport in Caco-2 cells
et al., 2002; Katneni et al., 2007; and P388/MDR cells
Mudra and Borchardt, 2010; In vivo 1–10% (w/v) Inhibition of epirubicin, digoxin,
Rege et al., 2002; Shono et al., rhodamine 123, and verapamil
2004; Yamagata et al., 2007a; (suggested) transport in situ rat
Yu et al., 2011a; Zhang et al., everted gut sacs
2003) Suggested no effect on digoxin transport in
isolated rat jejunal tissue
Tween 80 MRP-2 (Bogman et al., 2003) 0.0002–2% (w/v) No significant effect on methylfluorescein-
glutathione conjugate transport in
MDCK cells overexpressing MRP-2
BCRP (Yamagata et al., 2007a) 50–100 mM No significant effect on mitoxantrone
transport in BCRP-expressing cells
Glycerides and fatty acids
Capmul MCM P-glycoprotein and MRP-2 0.0002–2% (w/v) No significant effect on rhodamine 123
(Bogman et al., 2003) transport in P388/MDR cells or
methylfluorescein-glutathione conjugate
transport in MDCK cells
overexpressing MRP-2
Imwitor 742 P-glycoprotein (Cornaire et al., 0.05–0.5% (w/v) Inhibition of digoxin and celiprolol efflux
2004) in rat everted gut segments
Increased early plasma concentrations of
digoxin in rats
Miglyol P-glycoprotein (Cornaire et al., 0.05–0.5% (w/v) Inhibition of digoxin efflux rat everted gut
2004) segments. Little effect in vivo in rats
1-monoolein and P-glycoprotein (Barta et al., 500 and 100 mM Inhibition of rhodamine 123 transport in
1-monostearin 2008) Caco-2 cells
Peceol (glycerol P-glycoprotein (Risovic et al., In vitro 0.1–1% (w/v) Inhibition in rhodamine 123 transport in
monooleate) 2003, 2004; Sachs-Barrable Caco-2 cells
et al., 2008) In vivo 1 ml Enhanced amphotericin absorption in
vivo, due to increase solubilization
and permeability
Endogenous Solubilizing Species
Lecithin P-glycoprotein and MRP-2 0.0002–2% (w/v) No significant effect on rhodamine
(Bogman et al., 2003) 123 transport by P-gp or
methylfluorescein-glutathione
conjugate transport by MRP-2
Phospholipids P-glycoprotein (Lo, 2000) Phospholipid Inhibition of epirubicin transport in
liposomes Caco-2 cells and increased
absorption of epirubicin in
isolated rat jejunum and ileum
Sodium taurocholate P-glycoprotein (Ingels et al., FaSSIF Inhibition of talinolol, digoxin and
2004) doxorubicin transport in vitro
Caco-2 cells and in situ everted
rat jejunum and ileum
Sodium taurocholate P-glycoprotein (Ingels et al., 0–3 mM Inhibition of cyclosporine transport
2002) in Caco-2 cells
MDCK, Madin-Darby canine kidney.

solubilization (Aungst, 2000) appears to be required to Table 27). Several recent reviews have also addressed
enhance permeability, and there may be little differ- excipient-based transporter inhibition (Seelig and
ence in the concentrations required for permeability Gerebtzoff, 2006; Constantinides and Wasan, 2007;
enhancement versus those that result in cytotoxicity Werle, 2008; Li-Blatter et al., 2009; Goole et al., 2010).
(Seelig and Gerebtzoff, 2006). The proposed mechanisms by which lipidic excipients
More recently, an increasing number of studies have influence efflux activity include direct interaction with
described the ability of LBF components to enhance the transporter (Friche et al., 1990a; Spoelstra et al.,
drug permeability via inhibition of the major intes- 1991; Zordan-Nudo et al., 1993; Orlowski et al., 1998),
tinal efflux transporters P-gp, BCRP, and MRP (see changes to membrane fluidity or the microenvironment
Strategies for Low Drug Solubility 457

of membrane lipid domains (leading to indirect effects the GI fluids, the large surface area of the GI tract,
on transporter activity via protein destabilization) differences in dissolution rate of the drug and excipi-
(Woodcock et al., 1992; Loe and Sharom, 1993; Regev ent, changes to intestinal pH and motility, and
et al., 1999), or changes to efflux transporter expression intersubject and intrasubject variability will conspire
(Risovic et al., 2003, 2004; Sachs-Barrable et al., 2007; to limit excipient accumulation at the site of drug
Barta et al., 2008). absorption. Increases in permeability may also be
Seelig and coworkers suggest that surfactants that difficult to achieve in vivo as drug absorption and
are able to anchor into the membrane most effectively excipient effects on intestinal permeability are often
(via the presence of a hydrophobic tail) and that also site dependent. For example, drugs are commonly ab-
display P-gp binding motifs, for example, 1,2, dicaprylin, sorbed from the proximal small intestine, whereas
tricaprylin (Miglyol 808; Sasol, Hamburg, Germany), some permeation enhancers are more effective at
PEG oleate, Cremophor EL, Solutol HS-15, vitamin E, promoting passive diffusion in the distal intestine
vitamin E acetate, vitamin E TPGS, Tween 80, octyl- (Yamamoto et al., 1994; Yeh et al., 1994; Sekine et al.,
b-glucoside, and Triton X-100, are most capable of 1985; Aungst, 2000).
enhancing drug permeability in vivo. In contrast, the Finally, membrane properties and the expression
same authors suggest that high-molecular-weight exci- and activity of efflux transporters differ between in
pients, such as polyethylene glycols and polyoxyethylene- vitro and in vivo models. For instance, the intestinal
polyoxypropylene block copolymers (poloxamers), may membrane is seemingly more resistant to enhance-
be less likely to inhibit P-gp in vivo (despite in vitro ments in passive diffusion than in vitro models,
evidence to the contrary) owing to difficulties in possibly as a result of differences in the types of lipids
incorporation into the intestinal membrane, since the in the membrane and the greater lateral packing
intestinal membrane has a greater lateral packing density of the intestinal membrane compared with
density than most in vitro cell lines (Seelig and many cultured cell lines (Aungst, 2000; Seelig and
Gerebtzoff, 2006). They also suggest that further data Gerebtzoff, 2006). The intestinal membrane is also
are required to support whether membrane fluidiza- covered in a thick layer of mucus that is absent in most
tion is relevant to the in vivo mechanism of P-gp cell lines, and many in vitro cell lines, particularly
inhibition for most surfactants. For example, Rege immortalized cell lines, overexpress a number of the
et al. (2002) concluded that membrane fluidization is common transport proteins. The intestinal membrane
an unlikely mechanism of permeability enhancement may therefore be less readily accessed or less readily
as Tween 80 and Cremophor EL fluidize lipid bilayers, penetrated by surfactants that might otherwise
whereas vitamin E TPGS rigidifies rather than fluid- permeabilize/solubilize cell membranes or inhibit
izes lipid bilayers, yet all inhibit multiple efflux transporter efflux in vitro. Nonetheless, in vitro studies
transporters proficiently. provide a useful tool to generate rank-order relation-
a. Correlation of In Vitro Effects of Lipid-Based ships for permeability enhancing effects (Quan et al.,
Formulation Excipients on Permeability with Changes 1998; Aungst, 2000) and allow for greater insight into
to In Vivo Exposure? Although several studies have the mechanisms of permeation enhancement.
shown an effect of lipidic excipients on permeability in Despite the complexities of in vitro–in vivo correla-
vitro, reports of the effects on the permeability of poorly tion of permeability data, a number of in vivo studies
water-soluble drugs in vivo are more limited. In part, have shown enhanced absorption of hydrophilic or
this reflects the difficulty in delineating permeability macromolecular drugs after coadministration with
enhancement from solubilization since most of the lipidic excipients, including FA (Burcham et al., 1995;
amphiphiles that affect permeability also impact on Constantinides et al., 1996), monoglycerides (Beskid
solubilization. For poorly water-soluble drugs where et al., 1988; Constantinides et al., 1994; Hastewell et al.,
permeability and solubility limit bioavailability, ascrib- 1994; Lundin et al., 1997), and surfactants (Prasad
ing an increase in bioavailability to formulation effects et al., 2003; Rama Prasad et al., 2003), suggesting that
on either permeability or solubility is therefore complex. lipidic excipients may enhance the absorption of poorly
Studies using probes where solubility is less likely to permeable (but water-soluble) molecules via increases
limit bioavailability (e.g., BCS class 3) provides for easier in passive permeability.
data interpretation and may allow effects of the same In vivo studies in which lipidic excipients have been
surfactant to be inferred in studies with drugs where proposed to enhance oral absorption via inhibition of
solubility is a confounding issue. intestinal efflux include studies with both hydrophilic
Surfactant effects on permeability are likely to and poorly water-soluble, lipophilic drugs. These in-
depend on high localized concentrations of surfactant, clude talinolol (Bogman et al., 2005), saquinavir
regardless of whether the change to permeability ref- (Martin-Facklam et al., 2002a), amprenavir (Brouwers
lects inhibition of an efflux transporter or effects on et al., 2006), digoxin (Tayrouz et al., 2003), cyclosporine
passive transport. In vivo, however, this may be A (Sokol et al., 1991; Chang et al., 1996; Pan et al.,
difficult to achieve since dilution of the formulation in 1996), and paclitaxel (Varma and Panchagnula, 2005;
458 Williams et al.

Constantinides and Wasan, 2007) (all of which are dependent and include the potential for both direct
substrates for P-gp efflux) and topotecan (a substrate effects on enzyme activity and indirect effects via
for BCRP) (Yamagata et al., 2007a). The latter changes to enzyme expression levels (Patel and Brocks,
(Yamagata et al., 2007a), provides an excellent exam- 2009). For example, in hepatic microsomes, FAs di-
ple of a study design that has allowed differentiation of rectly inhibit Cyp isoenzymes, and the efficiency of
effects on permeability versus solubility for a poorly inhibition decreases in the following order: polyun-
water-soluble drug. In this work, the authors showed saturated FA . unsaturated FA . saturated FA
increases in exposure after coadministration of topotecan (Hirunpanich et al., 2007). In contrast, the expression
with Tween 85 and Pluronic P85 to mice but went on to levels of Cyp isoenzymes are either increased or
show no effects after i.v. administration (ruling out decreased on in vitro incubation with LC FA in rat
surfactant effects on systemic disposition) or after oral hepatocytes (Li et al., 2006, 2007) and after long-term
administration to BCRP knockout mice (ruling out ingestion of certain dietary oils in rats (Yoo et al., 1991;
effects on solubilization). Li et al., 1992; Brunner and Bai, 2000). Lipid
Coadministration of dietary lipids and LBF compo- coadministration can also influence systemic metabo-
nents may also affect presystemic metabolism. Exam- lism by altering systemic lipoprotein levels and
ples of increases in drug exposure as a result of therefore the nature of drug binding in the blood.
changes in presystemic metabolism stimulated by Association with lipoproteins has been shown to both
lipids (Patel and Brocks, 2009) or other solubilizing reduce and enhance drug uptake into metabolic cells
excipients (Buggins et al., 2007) have been reviewed and thus increase or decrease systemic metabolism
recently. The potential for lipids to influence drug (Wasan et al., 2008; Patel and Brocks, 2009). Decreases
exposure via effects on metabolism is well illustrated in free concentrations of halofantrine resulting from
by studies that show either increased (Gupta et al., increased binding to plasma lipoproteins have also
1990; Gupta and Benet, 1990) or decreased metabolism been suggested to reduce systemic clearance and vol-
(Liedholm and Melander, 1986; Milton et al., 1989; ume of distribution (Humberstone et al., 1998). In
Humberstone et al., 1996; Humberstone et al., 1998; contrast, increased lipoprotein binding has been
Meng et al., 2001) after drug administration to humans reported to increase, decrease, or not change the area
(Liedholm and Melander, 1986; Milton et al., 1989; under the plasma concentration versus the time profile
Gupta et al., 1990; Gupta and Benet, 1990; Meng et al., of cyclosporine, depending on the model of hyperlipid-
2001) or beagle dogs (Humberstone et al., 1996, 1998) emia used (Wasan et al., 2008). It seems unlikely,
with a high-fat meal. Similar effects have been however, that the small lipid doses contained in LBF
described in rats after administration with relatively will lead to changes in systemic lipoprotein concen-
large lipid doses (on a milligrams per kilogram basis). trations that are sufficient to influence drug metabo-
For example, coadministration with 2 ml/kg peanut oil lism via an effect on binding.
containing 1% cholesterol decreased the metabolism of In addition to lipids, excipients, including surfac-
halofantrine and amiodarone (Brocks and Wasan, tants and cosolvents, also influence drug metabolism in
2002; Shayeganpour et al., 2005) and 50–200 mg/kg vitro and in situ. Tween 20, Tween 80, vitamin E
docosahexaenoic acid appeared to reduce the metabo- TPGS, Cremophor EL, Cremophor RH40, Brij 35,
lism of cyclosporine, saquinavir, and midazolam by re- Solutol HS-15, and other nonionic surfactants inhibit
ducing intestinal but not hepatic metabolism (Hirunpanich Cyp3A4 and Cyp2C9 in a concentration-dependent
et al., 2006, 2008; Hirunpanich and Sato, 2006). Me- manner in liver microsomes or isolated hepatocytes at
tabolism of amiodarone was also reduced after in- concentrations below the CMC (Bravo González et al.,
cubation with intestinal segments isolated from rats 2004; Randall et al., 2011; Rao et al., 2010; Christiansen
preadministered 2 ml/kg peanut oil with 1% cholesterol et al., 2011). PEG400 and Pluronic 85 inhibit verapa-
compared with control rats (Shayeganpour et al., 2008). mil metabolism in isolated rat intestinal segments,
The metabolism of halofantrine by isolated intestinal whereas in the same model, TPGS showed only mar-
segments was similarly reduced in the presence of lipids ginal effects (Johnson et al., 2002). In an in situ rat
(Patel et al., 2010), and a trend toward lower systemic perfusion model PEG-400, but not Cremophor EL,
metabolite to parent ratios was evident after adminis- TPGS, or Tween 80 reduced verapamil metabolism
tration of anethol trithione to rats with lipid formula- (Mudra and Borchardt, 2010). Mountfield et al. (2000)
tions (Yu et al., 2011b). These studies demonstrate that further reported that a number of potential LBF com-
the glycerides and FA present in LBF can influence ponents, including oils, cosolvents, and surfactants,
drug metabolism. However, further studies are required were able to inhibit Cyp3A4 at relatively high concen-
to confirm that this is the case after in vivo adminis- trations, but they concluded that only the amphiphilic
tration of the smaller quantities of lipid typically con- excipients (Tween 80 and oleic acid) were likely to be
tained in LBF. capable of significant enzyme inhibition at the con-
The mechanism by which lipids are able to perturb centrations one would expect to find in vivo. Fewer
presystemic metabolism are drug and formulation studies have attempted to examine the potential for
Strategies for Low Drug Solubility 459

surfactants or cosolvents to inhibit Cyp-450-mediated efflux is believed to enhance the time available for
metabolism in vivo. Ren et al. (2009) demonstrated enterocyte-based presystemic metabolism. As such,
inhibition of midazolam metabolism in vitro and some inhibition of efflux might be expected to have the addi-
changes to in vivo metabolism of midazolam after tional benefit of providing protection from enterocyte-
administration with polysorbate 20, Cremophor EL, based metabolism by reducing the time available for
Myrj 52, and Pluronic F68. Inhibition of metabolism, in metabolism.
addition to effects on intestinal transporters, has also 4. Promotion of Lymphatic Drug Transport. In
been suggested as the mechanism by which formula- addition to potential effects on drug solubilization,
tions containing TPGS and Cremophor EL enhance the permeability, and metabolism, for a small subset of
absorption of cyclosporine (Chang et al., 1996) and highly lipophilic drugs, coadministration with lipids
saquinavir (Martin-Facklam et al., 2002a), respectively. can affect the route of drug trafficking to the systemic
How surfactants and cosolvents inhibit metabolism circulation through the promotion of drug transport via
and whether they are able to provide useful increases the intestinal lymphatic system. Lymphatic drug trans-
in exposure by inhibiting drug metabolism in vivo port has been described for many highly lipophilic drugs
remain unclear. Proposed mechanisms of inhibition in- and has been the subject of several reviews (Porter and
clude indirect effects on metabolic enzyme activity Charman, 2001b; O’Driscoll, 2002; Porter et al., 2007;
via membrane fluidization or depletion of ATP (as Trevaskis et al., 2008; Yanez et al., 2011).
has been described for P-glycoprotein mediated efflux The process of lymphatic drug transport is summa-
(Christiansen et al., 2011) or direct effects on enzyme rized in Fig. 64. After oral administration, drug
function or expression (Rao et al., 2010; Christiansen absorption typically involves uptake into and passage
et al., 2011). A number of surfactants are also di- across enterocytes into the underlying lamina propria,
gestible in vivo (Fernandez et al., 2007, 2008; Cuiné where both blood and lymphatic capillaries are pres-
et al., 2008; Christiansen et al., 2010). Digestion is ent. Blood drains from the mesenteric capillaries into
likely to alter significantly the physicochemical prop- the portal vein and subsequently flows via the liver to
erties of surfactants and, therefore, the potential for the systemic circulation. Blood flow from the intestine
inhibition of metabolic or efflux processes. This point, via the portal vein is some 500-fold faster than lymph
however, has not been addressed directly. In many flow through the mesenteric and thoracic lymph ducts;
cases, surfactant digestion will also liberate FA and the as such, the predominant drug absorption route for
liberated FA might also be expected to impact enzyme most drug molecules is via the blood.
activity, although the quantity of FA produced via As described in Section XI.A.1, ingestion of dietary fat
surfactant digestion will be limited. In addition to promotes the assembly of triglyceride-rich lipoproteins
digestion, the potential for surfactant degradation in (TRL), including chylomicrons and VLDL in the enterocyte,
the acidic conditions in the stomach, dilution in the GI and the subsequent transport of these lipoproteins
fluids, and poor absorption into enterocytes further from the intestine to the systemic circulation occurs via
complicate the likelihood of useful in vivo activity and the intestinal lymph (Kindel et al., 2010; Mansbach
in vitro-in vivo correlation. and Siddiqi, 2010). For some highly lipophilic drugs,
The impact of excipients on efflux and metabolism is association with these lipoprotein assembly and trans-
potentially linked by the proposed “metabolic alliance” port pathways leads to drug transport to the systemic
between efflux transporters and metabolic enzymes in circulation via the lymph rather than via the blood,
the GI tract (e.g., P-glycoprotein and Cyp3A4) (Benet and under some circumstances, notably after post-
and Cummins, 2001; Benet et al., 2004). In this model, prandial administration (where lipid absorption and

Fig. 64. Schematic of lymphatic drug transport. After oral administration, most drugs (D) are taken up into and pass across the enterocyte, access the
lamina propria and are then taken up into blood capillaries that converge into the portal vein and pass through the liver before accessing the systemic
circulation. In contrast, highly lipophilic (log D . 5) and lipid-soluble [triglyceride (TG) solubility .50 mg/g] drugs may associate with lipid absorption
pathways, and in particular with lipoproteins (LPs) that are formed after resynthesis of lipid digestion products such as fatty acids (FA) and
monoglycerides (MG) to TG. After exocytosis from the enterocyte, drug-lipoprotein complexes are taken up specifically into the intestinal lymphatics,
which converge to form the thoracic lymph duct. Thoracic lymph empties directly into the systemic circulation and, as such, transport via the lymphatic
system avoids passage through the liver and the potential for first-pass hepatic metabolism.
460 Williams et al.

lipoprotein synthesis are high), lymphatic transport


may be the dominant means of drug transport (Khoo
et al., 2001; Trevaskis et al., 2008). Since lymph drains
from the lymphatic capillaries into the mesenteric and
thoracic lymph ducts before emptying directly into the
systemic circulation, intestinal lymphatic drug trans-
port avoids first-pass hepatic metabolism.
In addition to dietary lipids, formulation-derived
lipids also stimulate TRL assembly and lymphatic
drug transport. In general, long-chain and monounsat-
urated glycerides stimulate lipoprotein formation more
effectively than medium-chain and saturated glycer-
ides and more effectively promote lymphatic drug
transport (Charman and Stella, 1986; Caliph et al.,
2000; Khoo et al., 2003). TRL production and lymphatic
drug transport are predictably increased by the admin-
istration of higher lipid loads (Charman and Stella,
1986; Khoo et al., 2003; White et al., 2009).
Historically, the physicochemical properties that
have been ascribed to drugs that are amenable to
significant lymphatic transport are a log P . 5 (or log D
for ionizable compounds) and solubility in long-chain
triglycerides .50 mg/g (Charman and Stella, 1986;
Porter and Charman, 2001b; O’Driscoll, 2002; Porter
et al., 2007; Trevaskis et al., 2008). These properties
reflect the necessity for avid drug partition into TRL to
drive drug transport via the lymphatics. Most lym-
phatically transported drugs conform to these charac-
teristics; however, more recent studies have also
suggested that in some cases, drugs with lower lipid
solubility but a specific affinity for lipoproteins beyond
simple partitioning behavior (e.g., affinity for the
lipoprotein interface) may also be significantly trans-
ported via the lymph (Gershkovich and Hoffman, 2005;
Trevaskis et al., 2010a; Gershkovich et al., 2009).
Nonetheless, all these compounds remain highly lipo-
philic as defined by log P.
For almost all lymphatically transported drugs,
coadministration with lipids leads to increases in
bioavailability and increases in lymphatic transport.
This has led to the suggestion that stimulation of
lymphatic drug transport is able to increase drug Fig. 65. (A) Systemic plasma concentration-time profiles after oral
absorption. Definitive evidence to support this conten- administration of 15 mg of CRA13 as a solidified micellar formulation to
tion, however, is lacking since delineation of lipid fed and fasted human volunteers. Values are expressed as means (fasted;
n 5 12, fed; n 5 6) 6 SEM. (B) Extent of absorption (white bar) and
effects on solubilization versus lymphatic transport is bioavailability (gray bar) of CRA13 in fasted noncannulated dogs (n 5 2)
difficult. Nonetheless, promotion of lymphatic drug and postprandial (fed) lymph-cannulated dogs (n 5 3) and the cumulative
transport of CRA13 in lymph over 10 h (black bar) in postprandial (fed)
transport can increase the oral bioavailability of drugs dogs (n 5 3), all expressed as a percentage of the administered dose. (C)
that are subject to significant first-pass metabolism Cumulative percent of the dose of CRA13 transported into the lymph over
time in postprandial lymph-cannulated dogs. Values are expressed as
(Khoo et al., 2001; Shackleford et al., 2003; Trevaskis means (n 5 3) 6 SD. Adapted from Trevaskis et al. (2009).
et al., 2009; White et al., 2009). For example, food has
a substantial positive effect on the oral bioavailability
of a lipophilic cannabinoid receptor agonist CRA13 in
humans (Fig. 65A). To investigate the mechanism by
which bioavailability is enhanced by food, the extent of nonlabeled CRA13 after a meal (Trevaskis et al., 2009).
absorption and bioavailability of CRA13 was compared Consistent with the data in humans, the absolute bio-
in fasted beagles administered radiolabeled CRA13 availability of CRA13 was low in fasted dogs (8–20%).
and in lymph-cannulated greyhound dogs administered However, this was not due to poor absorption as
Strategies for Low Drug Solubility 461

72–75% of radiolabeled CRA13 was recovered in the of poorly water-soluble drugs (Constantinides et al.,
systemic circulation (Fig. 65B), suggesting significant 2008). In general, parenteral LBF have been used for
first-pass metabolism. Bioavailability was increased to two primary purposes. First, they have been used as
47.5% in the dogs after postprandial administration a means to allow the delivery of lipophilic, poorly
(Fig. 65B), and most of the dose (43.7%) was trans- water-soluble drugs in a vehicle designed to be an-
ported to the systemic circulation via the lymphatic alogous to a cosolvent formulation or cyclodextrin
system (Fig. 65C). The total extent of absorption of solution (i.e., one that has little or no effect on
CRA13 in the fed lymph-cannulated dogs was estimated systemic pharmacokinetic parameters). In contrast,
at 63.2% (from addition of the extent of lymphatically many of the more complex parenteral LBF—such as
transported drug and the systemic availability of drug liposomes, solid lipid nanoparticles, certain micellar
in the blood, assuming 80% first-pass metabolism) (Fig. solutions, and various liquid crystalline phases—
65B). As CRA-13 was well absorbed in both fasted and have been used to intentionally influence pharmaco-
fed dogs, the positive effect of food on the bioavail- kinetic and biodistribution profiles and to target
ability of CRA13 appeared to reflect stimulation of drug delivery to specific anatomic or pathologic re-
lymphatic transport resulting in reduced first-pass gions (e.g., tumors) (Shi and Li, 2005; Constantinides
metabolism. et al., 2008). For example, liposomal formulations of
For drugs with very high first-pass metabolism, amphotericin (Janknegt et al., 1992) and doxorubicin
lymphatic transport can be responsible for the delivery (Lasic, 1996) have significantly improved patient
of most of the bioavailable drug to the systemic therapy by modifying pharmacokinetics to improve
circulation, even when the overall extent of lymphatic activity and safety profiles. The formulation of lipid-
transport is relatively low. For example, the oral bio- based parenteral formulations as targeted drug
availability of testosterone and methylnortestosterone delivery systems to optimize drug distribution pat-
is essentially zero owing to very high first-pass me- terns to sites of activity versus toxicity is beyond the
tabolism. The synthesis of lipophilic prodrugs of both scope of the current review. The interested reader is
compounds, however, increases the proportion of the directed to the following excellent reviews for further
dose that partitions into developing lymph lipoproteins information regarding liposomes: Lian and Ho
and therefore facilitates lymphatic drug transport and (2001), Constantinides et al. (2008), and Puri et al.
the direct delivery of bioavailable drug to the systemic (2009); and the following reviews are recommended
circulation. Although these prodrugs are relatively in- for solid lipid nanoparticles: Müller et al. (2000),
efficient, the degree of exposure is sufficient to support Mehnert and Mader (2001), Wissing et al. (2004),
an oral testosterone drug product (Andriol, Merck) Joshi and Muller (2009), Puri et al. (2009), and
(Shackleford et al., 2003; White et al., 2009). Recent Bunjes (2010).
evidence further suggests that lymphatic transport In contrast, parenteral lipid solutions or suspensions
may provide some protection from enterocyte-based (Chien, 1981; Zuidema et al., 1994; Murdam and
metabolism by sequestering drug into developing lip- Florence, 2000; Larsen et al., 2009; Weng Larsen and
oproteins (Trevaskis et al., 2006). Larsen, 2009) are simple to formulate and are com-
monly marketed for i.m., s.c. (predominantly for
B. Design and Formulation of Lipid-Based Formulations veterinary use), and intra-articular (Larsen et al.,
LBF comprise a diverse group of materials—in- 2008b) administration of lipophilic drugs. For example,
cluding lipid solutions, emulsions, microemulsions, lipid solutions of lipophilic prodrugs have been used
nanoemulsions, micellar solutions, liposomes, lipid successfully in the treatment of schizophrenia and
nanoparticles, and emulsion preconcentrates (anhy- hormone replacement therapy for more than 30 years
drous formulations that spontaneously emulsify on (Davis et al., 1994; Weng Larsen and Larsen, 2009;
contact with aqueous media and are commonly re- Leucht et al., 2011). The aim of these formulations is to
ferred to as self-emulsifying drug delivery systems or sustain drug release by providing a “depot” at the site
SEDDS). LBF may be liquid, semisolid, or solid at room of administration. The release of lipophilic drugs from
temperature and intended for oral or parenteral the lipid depot is slowed by partitioning behavior that
administration. favors drug association with the vehicle rather than
1. Lipid-Based Fomulations for Parenteral Admin- the surrounding interstitial fluid. Validation of the
istration. A number of different types of formulations reproducibility of drug release is therefore a critical
containing lipids—including lipid solutions, emul- design aspect to ensure against dose dumping (Larsen
sions, microemulsions, micellar solutions, liposomes, et al., 2009; Weng Larsen and Larsen, 2009). Long-
and solid lipid nanoparticles [including variations acting parenteral lipid solutions typically consist of
such as nanostructured lipid carriers (NLCs) (see a lipophilic drug (or lipophilic prodrug derivative)
Section XII.B) and lipid-drug conjugate (LDC) nano- dissolved in vegetable oil (e.g., sesame oil, soybean
particles]—have been used to facilitate effective i.v., oil, safflower oil, Miglyols, castor oil, cottonseed oil)
s.c., i.m., or local (e.g., intra-articular) administration (Strickley, 2004). In some cases, a solubilizing agent
462 Williams et al.

(e.g., benzyl benzoate) or antioxidant (e.g., tocopherol) and MC triglycerides typically exhibit greater oxida-
may be added. tive stability than LC triglycerides (as a result of the
Conventional oil-in-water emulsions and nanoemul- presence of fewer unsaturated fatty acid chains). The
sions are also widely used for parenteral delivery of choice of emulsifier is driven by toxicity, the need for
poorly water-soluble drugs during preclinical (Li and all components in parenteral formulations to be
Zhao, 2007; Shah and Agnihotri, 2011) and clinical metabolized or excreted, the potential for effective
evaluation (Floyd, 1999; Driscoll, 2006b; Hippalgaonkar emulsion stabilization, and the need for stability
et al., 2010). Intravenous lipid emulsions have been during sterilization. Natural lecithin, from either
used as a part of parenteral nutrition programs for animal (egg yolk) or vegetable (soybean) origin, has
more than 50 years: Intralipid (10% or 20% soybean oil, been used as an emulsifier in almost all marketed
1.2% egg phosphatides, and 2.5% glycerol) was in- products. Lecithin-stabilized emulsions require high-
troduced in 1961 and remains in use today (Floyd, pressure colloid milling to facilitate emulsification, but
1999; Driscoll, 2006a; Hippalgaonkar et al., 2010). the resulting oil-in-water dispersions are remarkably
Lipid emulsions have subsequently been used to facil- stable owing to strong adsorption of the lecithin at the
itate the delivery of lipophilic drugs via various oil-water interface. The aqueous phase of parenteral
parenteral administration routes, including i.v., s.c., emulsions contains ionic or osmotic agents, antioxidants,
i.m., and intra-articular administration. Indeed, several buffers, and preservatives (Floyd, 1999; Hippalgaonkar
marketed products use lipid emulsions to facilitate i.v. et al., 2010). Oils exert no osmotic effects, so isotonic
administration—including clevidipine, diazepam, propofol, adjustment (to 280–300 mosm/kg) is typically achieved
etomidate, perfluorodecalin, perfluorotripropylamine, via the addition of glycerol, sorbitol, or xylitol to com-
alprostadil, dexamethasone palmitate, flurbiprofen mercial preparations. Buffering agents are generally not
axetil, and vitamins A, D2, E, and K1 (Strickley, used as they may catalyze hydrolysis of TG and PL and
2004; Hippalgaonkar et al., 2010). The composition of reduce emulsion stability. A small amount of sodium
current products and considerations in the develop- hydroxide is often added to adjust to a slightly alkaline
ment, manufacture and approval (including safety pH (;8.0) as emulsion pH decreases on sterilization and
issues) of lipid emulsions have been reviewed else- storage because of the small amounts of hydrolysis of TG
where (Floyd, 1999; Strickley, 2004; Cannon et al., and PL to free FA.
2008; Driscoll, 2006a; Date and Nagarsenker, 2008; Parenteral emulsions are typically formulated by
Bunjes, 2010; Hippalgaonkar et al., 2010; Mirtallo dissolving drug in the oil phase before emulsification
et al., 2010) and are briefly addressed in the following with the aqueous phase (i.e., de novo emulsion for-
section. mation) or via the addition of drug to preprepared
Lipid emulsions are most often used for drugs with commercial parenteral emulsions. The latter approach
very low water solubility, where more traditional is a relatively simple way to prepare poorly water-
parenteral formulation options (e.g., cosolvents, pH soluble drugs for i.v. administration in preclinical
control) are not possible, and where the lipid solubility studies. In this case, the drug is predissolved in a co-
of the drug molecule is high. Other advantages include solvent and slowly added to a commercial lipid emul-
the potential to reduce local pain and irritation, sion under conditions of agitation (e.g., ultrasonication
improved drug stability, and the availability of large- or homogenization). The de novo method is more
scale methods of production such as homogenization. commonly used for clinical formulations, but drug
Disadvantages include the inherent thermodynamic may also be incorporated into preprepared i.v. emul-
instability of lipid emulsions, the limited number of sions with the aid of a cosolvent, via direct incorpora-
lipids and emulsifiers that have been approved for tion of drugs, which are liquid at room temperature
parenteral use, and the possibility of low drug loading (e.g., propofol), or via direct incorporation of nano-
(Mirtallo et al., 2010). Parenteral administration in milled drug powders or nanocrystals followed by
a lipid emulsion can also alter clearance and biodis- homogenization. De novo preparation is achieved by
tribution profiles (Hippalgaonkar et al., 2010; Caliph first dissolving (usually with heating) all water-soluble
et al., 2012), and these effects require consideration in and oil-soluble ingredients in the aqueous or lipid
both the clinical and preclinical setting. phase, respectively. Emulsifiers are then added to
Parenteral lipid emulsions generally consist of either the oil or aqueous phase. The lipid phase is
10–30% triglyceride (long chain and/or medium chain), mixed with the aqueous phase under controlled condi-
emulsifiers, and an aqueous phase. LC triglycerides tions (temperature, agitation, rate of addition) to form
approved for clinical use include triolein, soybean oil, a coarse emulsion (droplet size ,20 mm). The droplet
safflower oil, sesame oil, and castor oil, whereas MC size of the coarse emulsion is subsequently reduced (to
triglycerides include fractionated castor oil, Miglyol a maximum mean droplet size ,500 nm) via homo-
810 and 812, Neobee M5 (Stepan Company, Deerfield, genization or microfluidization. The emulsion is ulti-
IL), and Captex 300. Drug solubility (mg/ml) in MC mately filtered to remove any large particulates.
triglycerides is usually higher than in LC triglycerides, Sterilization is achieved through aseptic preparation
Strategies for Low Drug Solubility 463

or terminal heat sterilization or filtration. Glass to measure). Perhaps the more flexible definition is to
containers are preferred as poorly water-soluble drugs use the generic description of SEDDS formulations and
and lipophilic components may adsorb to plastic and subsequently to quote a particle size to define more
because plastics are more permeable to oxygen and accurately the dispersion obtained. Regardless of de-
susceptible to leaching oil-soluble plasticizers. finition, the feasibility of solid dosage forms containing
Characterization of the resulting emulsion involves liquid LBF is evident in the availability of several
the examination of visual appearance (creaming, co- marketed products (Strickley, 2007).
alescence, oil separation, color change), chemical anal- Solid and semisolid LBF are also being developed
ysis (drug and excipient and degraded product content, with greater frequency (Vasanthavada and Serajuddin,
e.g., FA), droplet size distribution and zeta potential, 2007; Cole et al., 2008; Jannin et al., 2008; Tang et al.,
viscosity, pH, preservative tests, sterility, and pyrogen 2008; Shah and Serajuddin, 2012; Tan et al., 2012),
testing (Driscoll, 2006b). For preclinical studies, prep- and a number of lipid formulation excipients are solid
aration under as near as possible to aseptic conditions or semisolid at room temperature. Examples include
followed by filtration before administration is usually hydrogenated vegetable oils, saturated fatty acids, and
sufficient, along with verification of drug content after glycerides and semisynthetic glycerides based on
preparation and filtration. Mean droplet size is an hydrogenated vegetable oils (e.g., Gelucires). A more
important indicator of emulsion stability, and the exhaustive list of excipient melting point and physical
presence of larger droplets (.5 mm) is likely to lead state data are available in Gibson (2007). Solid and
to fat embolism and toxicity. The US Pharmacopeia semisolid lipid fill materials minimize the chance of
limits for lipid injectable emulsions include pH be- capsule leakage and incompatibilities on storage (Cole
tween 6.0 and 9.0, mean droplet size #500 nm, volume et al., 2008) and may provide for some degree of
proportion of larger fat globules (.5 mm) # 0.05%, and sustained release (Vasanthavada and Serajuddin,
free fatty acids #0.07 mEq/g (Driscoll, 2006b). 2007; Jannin et al., 2008; Tang et al., 2008). However,
2. Lipid-Based Formulations for Oral Admin- drugs are typically incorporated into solid or semisolid
istration. For oral administration, LBF may be liquid, lipidic vehicles by mixing under elevated temperature
semisolid, or solid formulations at room temperature before capsule filling, and drug or lipid may phase-
and comprise lipid solutions, emulsions, microemul- separate and crystallize when the vehicle solidifies.
sions, micellar solutions, emulsion preconcentrates, More stable systems have been described in which
liposomes, and lipid nanoparticles. Liquid formulations solid SEDDS are prepared by dissolving drug in a melt
are useful in the preclinical setting as they are re- containing SEDDS excipients and high-molecular-
latively quick to develop and can be gavaged at a range weight PEG or poloxamer (to promote solidification)
of doses (Li and Zhao, 2007; Shah and Agnihotri, 2011). followed by hot filling into capsules and cooling (Li
They may also be of benefit in select patient groups et al., 2009b; Shah and Serajuddin, 2012). A range of
(pediatrics or those with swallowing difficulties), and novel formulation techniques have also been used to
a limited number of liquid LBF have been marketed transform liquid or semisolid formulations into solid
(Strickley, 2004, 2007). particles (powders, granules, or pellets) that can be
In most clinical applications, however, solid dose filled into capsules or sachets or compressed into
forms are preferred. To facilitate this, liquid LBF such tablets. These techniques have been reviewed in detail
as lipid solutions and self-emulsifying drug delivery (Vasanthavada and Serajuddin, 2007; Jannin et al.,
systems (SEDDS) are filled into hard or soft gelatin 2008; Tang et al., 2008) and may be used to enhance
capsules. SEDDS form emulsions with particle sizes in drug solubility and to promote absorption or to control
the nanometer to micrometer size range, and the drug release. Commonly used techniques include spray
terminology that has been used to describe these sys- cooling, spray drying, adsorption onto solid carriers,
tems is complex. Initially, second-generation SEDDS, melt granulation, melt extrusion, high-pressure ho-
which dispersed to form optically clear emulsions with mogenization (to produce solid lipid nanoparticles or
low particle sizes (typified by the cyclosporine Neoral nanostructured lipid carriers), and a range of super-
formulation), were classified as self-microemulsifying critical fluid–based methods.
drug delivery systems (SMEDDS) with the under- In spray cooling, a melt is sprayed into a cooled cham-
standing that what was produced was a microemulsion. ber in which the molten droplets congeal and recrystallize
However, the definition of a microemulsion is thermo- into spherical solid particles. Polyoxylglycerides, par-
dynamic (i.e., microemulsions are thermodynamically ticularly stearoyl polyoxylglycerides (Gelucire 50/13)
stable, unlike all other emulsions) and does not relate (Cavallari et al., 2005; Passerini et al., 2006), are fre-
to particle size. Indeed, most pharmaceutically rele- quently used for spray-cooled products. In contrast,
vant microemulsions have particle sizes in the nano- spray drying involves spraying a liquid solution into
meter size range. This has led to the use of SNEDDS as a heated rather than a cooled chamber to evaporate
a descriptor based on particle size rather than the solvent (organic solvent or water), yielding solid
thermodynamic stability (which is inherently difficult microparticles. In this way, LBF excipients (e.g., lauroyl
464 Williams et al.

or stearoyl polyoxylglycerides) either alone or in combi- a gradual decrease in temperature and pressure to
nation with a solid carrier may be solubilized in organic promote phase separation and the generation of lipid-
solvent and spray dried to remove the solvent (Chauhan coated drug particles or solid dispersions containing
et al., 2005). Alternatively, oil-in-water emulsions may lipids. Lipid excipients that have been used during the
be spray dried to prepare dry emulsions. Medium-chain production of supercritical fluid–based SLNs include
triglycerides and oleyl polyoxylglycerides have been glyceryl trimyristate (Dynasan 114; Sasol), stearoyl
used as the lipophilic phase in dried emulsions of this polyoxylglycerides (Gelucire 50/13), vitamin E TPGS,
type (Christensen et al., 2001; Dollo et al., 2003; and lauroyl polyoxylglycerides (Gelucire 44/14) (Sethia
Hansen et al., 2004; Jang et al., 2006; Yi et al., 2008; and Squillante, 2002; Ribeiro Dos Santos et al., 2003;
Oh et al., 2011; Kang et al., 2012). To generate Thies et al., 2003). Techniques for formulating solid
adsorbed solid LBF, liquid lipid formulations such as dispersions containing lipids and SLN/NLC are de-
SEDDS are simply adsorbed to a carrier (e.g., calcium scribed in more detail in Sections X and XII.B,
silicate, magnesium aluminometasilicate, silicon di- respectively.
oxide, or carbon nanotube) under mixing to form a free- 3. The Lipid Formulation Classification System.
flowing powder (Ito et al., 2005; Patil and Paradkar, Pouton (2000) introduced the Lipid Formulation
2006; Abdalla et al., 2008; Agarwal et al., 2009; Tan Classification System (LFCS) in 2000 to describe and
et al., 2009; Dixit and Nagarsenker, 2010; Wang et al., classify LBF according to their composition and
2010c; Tan et al., 2011; Van Speybroeck et al., 2012). behavior on dispersion and digestion. The LFCS was
Solid or semisolid lipids (e.g., polyoxylglycerides, subsequently updated in 2006 to include an additional
lecithin, partial glycerides, polysorbates) (Seo et al., classification (Pouton, 2006). There are currently four
2003; Shimpi et al., 2005; Vilhelmsen et al., 2005) have classes of LBF listed in the LFCS. The advantage of the
also been used to create solid SEDDS via traditional LFCS lies in its simplicity; however, there are also
melt granulation, effectively forming an emulsifiable limitations when trying to classify all possible combi-
solid dispersion (see Section X). Melt granulation may nations of excipient into four categories and more
be used further to adsorb semisolid self-emulsifying inclusive, albeit more complex, classifications can be
systems onto solid carriers (Gupta et al., 2001, 2002). envisaged (Müllertz et al., 2010). The lipid composition
In contrast, melt extrusion techniques force molten of the current LFCS classifications and a comparison of
lipid-based formulations through a die under con- their behavior on dispersion and digestion are listed in
trolled conditions to produce a product that sub- Table 28.
sequently solidifies to form extrudates of uniform Type I LBF comprise drugs dissolved in a digestible
shape and density. Lipid-based excipients, such as oil and may contain triglycerides, diglycerides, mono-
sucrose monopalmitate (Surfhope D-1616; Halim Bio- glycerides, or mixtures thereof. Type I formulations are
tech, Singapore, Singapore), lauroyl polyoxylglycerides therefore most appropriate for drugs with high solu-
(Gelucire 44/14), and polysorbate 80 (Tween 80) have bility in simple glyceride lipids. Type I formulations
also been included as additives in traditional melt have the advantage of simplicity and are resistant to
extrusion formulations (e.g., those containing microcrys- precipitation on capsule rupture. The major complexity
talline cellulose) to enhance the dissolution of poorly of type I formulations is the requirement for digestion
water-soluble drugs (Hulsmann et al., 2000; Serratoni to promote formulation dispersion and the limited
et al., 2007). More complex extrusion techniques have solvent capacity for many drugs.
described the specific introduction of lipidic excipients Type II formulations are isotropic mixtures of lipids
(such as Gelucire 44/14) into the core of sustained and lipophilic surfactants (HLB , 12) that self-emulsify
release formulation matrices (Mehuys et al., 2004, to form 200 nm–1 mm particle size emulsions on contact
2005; Windbergs et al., 2010). with aqueous media (i.e., typical SEDDS behavior). Self-
Solid lipid nanoparticles (SLNs, which possess a solid emulsification occurs when the surfactant concentration
core) or nanostructured lipid carriers (NLCs, which exceeds 25% w/w, and the optimum concentration is
possess a liquid core) may be produced by high- often 30–40%. Viscous liquid crystalline phases may
pressure homogenization (Müller et al., 2000; Mehnert occur in formulations with higher surfactant concen-
and Mader, 2001; Hu et al., 2004c; Puri et al., 2009; trations (i.e., .50%). Type II systems have the advan-
Bunjes, 2010). SLNs typically consist of a solid matrix tage of low hydrophile content (and therefore a reduced
of high melting lipids such as glyceryl dibehenate likelihood of drug precipitation on dispersion), coupled
(Compritol 888 ATO; Gattefossé) and surfactants such with effective self-emulsification. Although some of the
as poloxamer 188 (Pluronic F68) or Tween 80. NLCs first publications to describe SEDDS formulations
generally contain a liquid lipid excipient such as MC centered around type II systems (Charman et al.,
triglycerides in addition to the components of SLNs. 1992), more recently the popularity of type III and type
Supercritical fluid–based methods have also been used IV LBF and the limited number of approved lipophilic
to generate SLNs where the drug and lipidic excipients surfactants that support self-emulsification have lim-
are dissolved in a supercritical fluid, followed by ited the application of type II systems.
Strategies for Low Drug Solubility 465
TABLE 28
Typical composition of LFCS types I, II, IIIA, IIIB, and IV lipid-based formulations and their solubilization capacity for drug before and after dispersion
and digestion in the gastrointestinal tract
Increasing Hydrophilic Content →

Typical
Composition Type I Type II Type IIIA Type IIIB Type IV
(% w/w)

Triglycerides or 100 40–80 40–80 ,20


mixed
glycerides
Water-insoluble 20–60 0–20
surfactants
(HLB ,12)
Water-soluble 20–40 20–50 30–80
surfactants
(HLB .12)
Hydrophilic 0–40 20–50 0–50
cosolvents
Solvent capacity Often poor but may be Generally . type I Generally . type II Generally . type IIIA High, usually greatest
prior to acceptable for highly of all types
dispersion lipophilic drugs
Solubilization Coarse emulsion Turbid emulsion of Clear, fine emulsion Clear, fine emulsion Very fine micellar
capacity upon particle size of particle size of particle size solution
dispersion 250–2000 nm 100–250 nm 50–100 nm
Limited precipitation Unlikely to lose May lose solvent Likely to lose some High risk of
solvent capacity capacity on solvent capacity precipitation on
on dispersion dispersion on dispersion dispersion
Impact of Digestion crucial for Digestion may not be Digestion perhaps Digestion unlikely to Probably not digested
digestion on absorption to critical for absorption not critical for be critical for extensively and not
solubilization proceed but likely to occur and absorption and may absorption, i.e., critical to
capacity and enhance dispersion be inhibited by drug absorption absorption.
absorption to smaller particle hydrophilic surfactants. without digestion Digestion restricted
size Digestion may reduce possible to surfactant
Possible loss of solvent Possible loss of solvent solubilization capacity Digestion may reduce hydrolysis but may
capacity on digestion capacity on digestion solubilization reduce solubilization
capacity capacity

Type III formulations include lipids, hydrophilic sur- III. The disadvantage of type IV formulations is that the
factants (HLB . 12), and cosolvents and self-emulsify high quantity of water-miscible excipients renders them
very effectively to form emulsions with small particle highly susceptible to drug precipitation on dispersion.
sizes (usually ,250 nm). Type III LBF have therefore Combinations of cosolvents and surfactants are used in
been described as self-microemulsifying drug delivery an attempt to reduce the extent of precipitation on
systems (SMEDDS) (based on thermodynamic expec- dilution of pure cosolvent solutions and to increase the
tations of the dispersion) or self-nanoemulsifying drug dispersion of pure surfactant solutions (which often gel
delivery systems (SNEDDS) (based on particle size). on contact with aqueous fluids). The lack of traditional
Type III formulations contain long- or medium-chain lipids in type IV formulations restricts the impact of
glycerides, high HLB surfactants (e.g., Cremophor EL, digestion on formulation performance; however, surfac-
Cremophor RH40, Tween 80, Gelucire), and cosolvents. tant digestion is likely and in some cases may alter
There is a somewhat arbitrary division between type formulation properties. Assessment of behavior under
IIIA and IIIB formulations to distinguish those that simulated dispersion and digestion conditions is there-
contain greater quantities of hydrophilic components fore recommended.
(type IIIB). Key considerations to formulation design 4. Selection of Lipid-Based Formulation Excipients.
include protection against drug precipitation on dis- Many of the lipidic excipients that are commonly used
persion and digestion of the formulation. In vitro in LBF have been tabulated in detail elsewhere
testing of the potential for precipitation on dispersion (Strickley, 2004; Gibson, 2007; Strickley and Oliyai,
and digestion is therefore crucial to development and 2007). The FDA maintains a list of excipients that have
assessment (see Sections XI.C.1 and XI.C.2). Most been previously approved and incorporated in mar-
currently marketed LBF are type III formulations. keted products (the Inactive Ingredient Guide, or IIG).
Type IV formulations were more recently added to the The IIG provides a useful database of approved excip-
LFCS (Pouton, 2006) and comprise mixtures of surfac- ients and the maximum current use levels by route of
tant and cosolvent in the absence of traditional lipids. administration (http://www.accessdata.fda.gov/scripts/
The attraction of type IV formulations is the high cder/iig/index.cfm). Table 29 gives a very brief sum-
solvent capacity of cosolvents and surfactant compared mary of the most common excipients employed in LBF.
with lipids, which usually permits higher drug loading A flowchart providing some general guidance for
in type IV formulations compared with types I, II, and excipient selection in the assembly of LBF is provided
466 Williams et al.

TABLE 29
Excipients commonly used in oral lipid-based formulations
Excipient Class Examples

Medium-chain triglycerides Coconut or palm seed oil, Miglyol, Captex


Long-chain triglycerides Soybean, sesame, safflower, sunflower, corn, cottonseed, olive, palm, peanut (arachis),
rapeseed oils, and hydrogenated vegetable and soybean oils
Medium- and/or long-chain monoglycerides, Medium chain: Capmul, Imwitor, Labrafac, Capmul GMO
diglycerides, or mixed glyceridesa Long chain: Maisine 35-1, Peceol, Geleol
Low HLB surfactants (HLB ,8)a Lipophilic sorbitan fatty acid esters (e.g., Span 85, Span 80), and esters of propylene
glycol and fatty alcohols (Lauroglycol, Capryol). Oleic acid
Water-insoluble, lipophilic surfactants (HLB 8-12) Predominantly oleate esters and include polysorbate 85 (Tween 85) and Tagat TO
Water-soluble surfactants (HLB . 12) Include products synthesized by reacting alcohols with ethylene oxide to produce
alkyl ether ethoxylates (e.g., Brij, cetomacrogol); sorbitan esters with ethylene
oxide to form sorbitan ester ethoxylates (e.g., Tween); ethylene oxide with castor
oil to produce castor oil ethoxylates (e.g., Cremophor EL, RH40); and vegetable
oils with PEG to produce mixtures of monoglycerides, diglycerides, and
triglycerides and PEG esters of fatty acid (Labrasol, Labrafil, Gelucire)
Cosolvents PEG 400, propylene glycol, propylene carbonate, ethanol
Lipid-soluble antioxidants a-Tocopherol, b-carotene, butylated hydroxytoluene (BHT), butylated
hydroxyanisole (BHA), propyl gallate
Polymeric precipitation inhibitors (PPIs) Most effective appear to be cellulose derivatives, e.g., hydroxypropyl
methylcellulose (HPMC), HPMC acetate succinate (HPMCAS)
Others include polyvinylpyrrolidone (PVP) and poloxamers
HLB, hydrophilic-lipophilic balance.
a
The distinction between polar oils and low HLB surfactants is difficult, and glycerides, especially monoglycerides of medium-chain lipids, might be classified as either.

in Scheme 4. Generally, LBF development is initiated plotted on three sides of a triangular phase diagram.
via a solubility screen to identify excipients capable of Immiscibility may not be apparent immediately, and
dissolving the required dose, followed by dispersion therefore physical stability should be assessed over
and digestion testing to identify combinations that time. Particular care should be taken when waxy
most readily prevent drug precipitation during GI excipients are heated before or during blending since
processing. Additional factors to consider when select- this may lead to supersaturation on cooling to room
ing excipients for LBF and in selecting the type of temperature. In general, type I and type IV LFCS
glyceride are summarized in Tables 30 and 31 and in formulations are most readily miscible, whereas type III
the sections that follow. Considerations surrounding systems containing LC oils and hydrophilic surfactants
encapsulation (Cole et al., 2008), stability (Cannon, or cosolvents present the most problems with respect to
2008), scale-up (Sirois, 2007), and regulatory approval miscibility. Type III systems usually require the
(Chen, 2008) of LBF have been reviewed and are not presence of polar oils, such as mixed glycerides, to
addressed here. promote mutual miscibility. An additional benefit of the
a. Solvent Capacity. With the exception of highly addition of mixed glycerides is the potential for these
lipophilic drug molecules, triglycerides are relatively more polar oils to promote formulation dispersion. In
poor solvents for poorly water-soluble drugs; as such, contrast to type III systems, type II systems are readily
most LBF also contain more polar oils, surfactants, or miscible as both MC triglyceride and LC triglyceride are
cosolvents. Care needs to be taken, however, to balance miscible with lipophilic surfactants. Mixed glycerides
the inclusion of hydrophilic excipients to increase are therefore not a requirement in type II systems, but
solvent capacity with the propensity for drugs to pre- they may be added to optimize performance.
cipitate on dispersion of formulations containing large c. Toxicity/Irritancy. In general, glycerides are
quantities of surfactants and cosolvents. To facilitate regarded as nontoxic as they are a common dietary
more effective measurement of solubility in lipidic component. Intravenous administration of large quan-
vehicles, in vitro solubility screening methods have tities (.2.5 g/kg/day) of lipid, however, can lead to
been described that are rapid; inexpensive; minimally excessive lipid load (Mirtallo et al., 2010). After oral
labor-intensive; amenable to high-throughput automa- administration, toxicity and irritancy are more likely
tion, parallel processing, and miniaturization; and to be issues with surfactants that may penetrate,
require only small quantities of drug (milligram or fluidize, or solubilize biologic membranes. As such,
submilligram) (Dai et al., 2008b). surfactant inclusion needs to be considered carefully,
b. Mutual Miscibility. Excipient miscibility in LBF particularly where products are expected to be used
is required to ensure stability and drug content chronically. Notable trends in surfactant toxicity in-
homogeneity. Ternary or pseudoternary phase diagrams clude toxicity of ionic surfactants typically . nonionic
are commonly used to identify regions in which surfactants; cationic . anionic surfactants; single-
excipients are miscible. These phase diagrams allow chain . bulky surfactants (e.g., polysorbates or poly-
easy representation of phase behavior and miscibility as ethoxylated vegetable oils) and ethers . esters (which
a function of the concentration of excipients in LBF are digestible).
Strategies for Low Drug Solubility 467

Scheme 4. A flowchart providing a general guide to lipid-based formulation design.

Nonionic surfactants are therefore more usually of a series of esterification products and a lack of ex-
included in products for oral use. After parenteral plicit molecular characterization of the product. Poten-
administration, toxicity is more likely due to direct tial batch-to-batch variation in surfactant properties
introduction into the systemic circulation. For this is therefore particularly evident for surfactants of this
reason, emulsifiers based on dietary lipids such as type where variations in the quantities of each product
lecithin are used in most parenteral products (see are likely as a function of batch.
Section XI.B.1). e. Capsule Compatibility. Compatibility of LBF
d. Purity and Chemical Complexity. Peroxides and with hard and soft gelatin capsules was reviewed in
aldehydes are present in trace quantities in many LBF detail by Cole et al. (2008). In general, low-molecular-
excipients and have the potential to impact negatively weight polar molecules such as cosolvents (e.g., pro-
on product stability. Excipients with low levels of trace pylene glycol), surfactants, and water are most likely to
contaminants are therefore preferred, and consistency interact with capsule shells, although differences are
in trace contaminant profiles between batches should evident between hard and soft gelatin capsules. Newer
be confirmed. Furthermore, many surfactants are ob- capsule materials such as hydroxypropyl methylcellu-
tained via hydrolysis and esterification of glycerides lose hard gelatin capsules and plant polysaccharide-
with PEG or ethylene oxide. This leads to the generation based soft gel capsules provide additional options with
468 Williams et al.

TABLE 30 TABLE 31
Factors affecting the choice of excipients for lipid-based formulations Comparative properties of glyceride lipids in oral lipid-
based formulations
Solvent capacity
Miscibility Triglycerides vs. monoglycerides, diglycerides, or mixed glycerides
Regulatory issues: irritancy, toxicity, knowledge, and experience Partial glycerides are often better solvents than triglycerides
Morphology at room temperature (i.e., melting temperature) (unless the drug is highly lipophilic), and their increased
Self-dispersibility and role in promoting self-dispersion of the amphiphilicity typically results in improved emulsification
formulation properties.
Digestibility,and fate of digested products Addition of mixed glycerides to combinations of triglyceride and
Capsule compatibility surfactants typically improves miscibility.
Purity, chemical stability Glyceride lipids comprising long-chain vs. medium-chain fatty acids
Cost of goods LC glycerides usually have lower solvent capacity and emulsify
less readily than MC glycerides.
LBF based on LC glycerides are often more effective at
maintaining solubilization capacity on dispersion and digestion.
respect to capsule compatibility and may provide LC glyceride more effectively promote lymphatic transport of highly
lipophilic drugs compared with MC glycerides.
advantage in some cases. Capsule manufacturers are LC lipids are typically less stable to oxidation.
an excellent source of information regarding excipient Unsaturated glycerides vs. saturated glycerides
compatibility and should be contacted for further Glycerides comprising unsaturated fatty acids usually exhibit
increased solvent capacity.
information. Unsaturated lipids are less stable to oxidation.
f. Polymers. Polymeric precipitation inhibitors Unsaturated glycerides more effectively promote lymphatic
transport of highly lipophilic drugs compared with saturated
(PPIs) may be included in formulations to reduce drug lipids.
precipitation during formulation dispersion and di- Saturated glycerides have higher melting points and are commonly
gestion in the GI tract (see Section XI.A.2) (Erlich solid at room temperature.

et al., 1999; Gao et al., 2004; Gao and Morozowich,


2006; Warren et al., 2010; Gao et al., 2011a). PPIs aim
to maintain the drug in a supersaturated, thermody- and precipitated drug via centrifugation, and analysis
namically unstable (metastable) state (where drug is of drug content in the precipitate or solubilized phase.
solubilized at concentrations above equilibrium solu- Formulations that disperse slowly or form coarse emul-
bility) for a period that is sufficient to facilitate sions can usually be identified by visual inspection,
absorption. Solubilization at supersaturated concen- although particle size can be quantified more specifi-
trations may also provide additional benefits via an cally with Fraunhofer diffraction sizers and photon
increase in thermodynamic activity. Polymers that are correlation spectrometers. Although complete disper-
commonly used in LBF are listed in Table 29. sion to form monodisperse colloidal systems provides
for rapid processing in vivo, particle size provides a
C. Assessment of Lipid-Based Formulations poor indicator of absolute in vivo performance since the
1. In Vitro Dispersion Testing. Dispersion testing is properties of most formulations are changed dramat-
conducted to identify formulations that disperse slowly ically by dilution and digestion before drug absorption.
or those that lead to drug precipitation during the For example, systems containing high proportions of
dispersion process. Many different methods have been surfactant or cosolvent commonly produce emulsions
described, but a typical dispersion test involves simple with very small particle sizes but may rapidly lose sol-
dilution (1:50, 1:100, or 1:250) of the formulation into vent capacity on further dilution or digestion lead-
biorelevant media (Jantratid and Dressman, 2009; Di ing to drug precipitation (Porter et al., 2004a; Sek
Maio and Carrier, 2011) and subsequent assessment of et al., 2006; Cuiné et al., 2007; Mohsin et al., 2009;
the extent of precipitation. The choice of media de- Prajapati et al., 2012). Assessment of solubilization
pends on whether behavior under gastric or intestinal behavior is therefore more critical than particle size
conditions or under fasted or postprandial conditions is characterization.
required. The formulation may be added dropwise as 2. In Vitro Digestion Models. In vitro digestion
a liquid or, preferably, where a solid dosage form is testing is used to examine the potential for drug pre-
anticipated, as a capsule placed in the media under cipitation from LBF during in vivo digestion. Further-
appropriate conditions of temperature and agitation. more, when coupled with biophysical methods (e.g.,
Dispersion testing can be conducted in standard dis- dielectric viscosity and conductivity measurements,
solution apparatus and using, for example, FaSSGF, scattering techniques such as dynamic light scattering
FaSSIF, or FeSSIF media (Dressman et al., 1998; (DLS), small-angle neutron scattering (SANS), or small-
Dressman et al., 2007). The difference between disper- angle X-ray scattering (SAXS) (Fatouros et al., 2007b),
sion testing for LBF and a standard dissolution test is NMR, raman spectroscopy, and electron microscopy
therefore the emphasis on maintenance of drug solu- (Fatouros et al., 2007a), electron paramagnetic reso-
bilization and detection of unwanted drug precipitation nance (EPR) (Abdalla and Mäder, 2009), multiplex
rather than the transfer of drug from the solid state coherent antistoked Raman scattering microspectro-
into solution. Drug precipitation can be assessed by sam- scopy (CARS) (Day et al., 2010) atomic force microscopy
pling as a function of time, separation of solubilized and cryo-TEM (Müllertz et al., 2012), the in vitro
Strategies for Low Drug Solubility 469

digestion model may also be useful in determining the oil phase, an aqueous phase containing mixed micelles
nature of the colloidal phases formed, as well as the fate and vesicles, and a pellet phase. Drug that precipitates
of a drug (in particular, crystallinity/amorphous char- during digestion is contained within the pellet phase
acter) (Sassene et al., 2010), during or after digestion of and is thought to represent drug that is poorly
different LBF. Several variations of the in vitro di- available for absorption since redissolution is required
gestion model have been described, and differences in and the dissolution of poorly water-soluble drugs from
experimental conditions are detailed elsewhere (Porter crystalline solids is poor. In contrast, drug that re-
et al., 2007, 2008; Dahan et al., 2008; Fatouros and mains solubilized within the aqueous phase is expected
Mullertz, 2008; Larsen et al., 2011). However, efforts to to be in rapid equilibrium with drug in free solution
develop standardized in vitro tests for LBF have led to
and to provide a reservoir of drug that is highly avail-
the Lipid Formulation Classification System (LFCS)
able for absorption. Using variations of this general
Consortium, which is an industry-academic consortium
protocol, several groups have demonstrated rank-order
that aims to identify consistent and appropriate in vitro
digestion conditions (Williams et al., 2012a,b). correlations between drug solubilization in the aqueous
Typical in vitro digestion conditions are described in phase on in vitro digestion and drug bioavailability in
Fig. 66. The conduct of an in vitro digestion test usually animals (Porter et al., 2004a,b; Dahan and Hoffman,
involves the dispersion of a LBF in temperature- 2006, 2007; Cuiné et al., 2007, 2008; Fatouros et al.,
controlled (37°C) media consisting of buffer, bile salt, 2008; Larsen et al., 2008a; Han et al., 2009; Tan et al.,
and phospholipid, followed by the initiation of diges- 2011; Ahmed et al., 2012). In vitro digestion testing has
tion via the addition of a source of pancreatic enzymes. also been reported to provide an indication of in vivo
Samples of the digest are then taken throughout the formulation behavior in humans (Taillardat et al.,
experiment and centrifuged to separate an undigested 2007).

Fig. 66. Experimental setup for an in vitro digestion experiment. (A) Lipid formulations are introduced into a temperature-controlled reaction vessel
(37°C), which contains media consisting of buffer and bile salt and phospholipid concentrations designed to simulate fasted or fed intestinal conditions.
The contents of the reaction vessel are stirred throughout experiments. Either at the time of introduction of the formulation or after a period of
dispersion (to enable assessment of drug precipitation upon dispersion), a source of pancreatic enzyme(s) is added to the system to initiate the process
of lipid digestion. Digestion of glycerides liberates FAs, which causes a transient drop in pH. The pH is maintained at a setpoint (B) by the addition of
NaOH via an autoburette coupled to a pH-stat meter controller. Maintenance of constant pH is important to mimic in vivo conditions and to enable
digestion to continue. The rate of NaOH addition is recorded via the computer. This enables the calculation of the extent of digestion as digestion of
a molecule of triglyceride liberates a molecule of monoglyceride and 2 free FAs and OH ions react with H ions from liberated free FA in a 1:1
stoichiometric ratio (assuming complete ionization). (C). Samples can be taken throughout the digestion process, digestion stopped via the addition of
the lipase inhibitor 4-bromophenylboronic acid (BPBA), and samples centrifuged to obtain three separate phases: an undigested oil phase, an aqueous
phase, and a pellet phase. Precipitated drug is present in the pellet phase, whereas drug solubilized in micellar structures within the aqueous phase is
believed to represent drug that is available for absorption. Adapted from Porter et al. (2007).
470 Williams et al.

Fig. 67. (A) Percent of drug mass contained within the aqueous phase (gray area of bars) and pellet phase (white area of bars). (B) Final aqueous phase
concentration of danazol after the addition of 250 mg of medium-chain (MC) or long-chain (LC) SEDDS formulation containing 15 mg/g danazol to an in
vitro digestion apparatus followed by dispersion for 30 min in nondigesting conditions or dispersion for 10 min followed by addition of pancreatin to
stimulate digestion for 30 min. Samples were ultracentrifuged to obtain an aqueous phase and pellet phase to probe for drug precipitation. (C) and (D)
Plasma concentration versus time profiles for danazol after oral administration of 15 mg of danazol to beagle dogs. Values are expressed as means (n 5
4) 6 SEM. (C) Profiles after fasted administration of lipidic solutions containing 5 mg/g of danazol in 3 g of MC-SEDDS, LC-SEDDS, or LC triglyceride
(soybean oil) solution. (D) Profiles following fasted or postprandial administration of a dry micronized powder form of danazol. The MC SEDDS
formulation consisted of the following ratios of components: danazol (5 mg), Captex 355 (360 mg), Capmul MCM (180 mg), Cremophor EL (360 mg), and
absolute ethanol (100 mg). The LC SEDDS formulation consisted of the following ratios of components: danazol (5 mg), soybean oil (300 mg), Maisine
35-1 (300 mg), Cremophor EL (300 mg) and absolute ethanol (100 mg). Adapted from (Porter et al., 2004a).

As an example, Fig. 67 shows data obtained after in SEDDS where exposure to danazol was significantly
vitro dispersion and digestion of MC SEDDS and LC greater after administration of the LC SEDDS
SEDDS containing danazol and the corresponding in compared with the MC SEDDS (Fig. 67C). Indeed,
vivo data obtained after fasted administration of the exposure after administration of the SEDDS was
same formulations to beagle dogs. Also shown are similar to that seen after postprandial administration
exposure profiles obtained after fasted administration of a traditional micronized powder formulation and
of a LC triglyceride solution of danazol and fasted and was significantly greater than that seen after
postprandial administration of a micronized powder administration of the micronized powder formulation
formulation of danazol to the beagles. Both the MC and of danazol to fasted beagles. These results demonstrate
LC SEDDS formulations supported solubilization of the utility of LBF in supporting exposure of lipophilic
danazol after in vitro dispersion under nondigesting poorly water-soluble drugs, the potential for LBF to
conditions, with ,12% of the drug present in the pellet overcome food effects, and the potential benefits of the
phase (Fig. 67A). However, after digestion, there was in vitro digestion model in identifying formulations
significant precipitation of danazol from the MC that are likely to perform poorly in vivo.
SEDDS, whereas danazol remained predominantly 3. In Vivo Studies. In vivo evaluation of the per-
(.95%) solubilized in the aqueous phase after 30 min formance of LBF is performed in a similar manner to
of in vitro digestion of the LC SEDDS (Fig. 67B). These that used for most other formulations. Unlike many
results were reflected in the in vivo performance of the traditional formulation approaches, however, the
Strategies for Low Drug Solubility 471

quantity of formulation administered (on a milliliters previously (Porter and Charman, 2001b,c; O’Driscoll,
per kilogram basis) is likely to play a significant role in 2002; Porter et al., 2007; Trevaskis et al., 2008).
ongoing drug solubilization and should therefore be In addition to assessment in lymph cannulated
controlled. Formulations may be administered via oral animal models, interest has increased in alternate
gavage (either directly or after dispersion in a small (and less surgically complex) animal models and in in
volume of water) or after encapsulation. Capsules can vitro predictive models. These include bioavailability
be hand-filled for large animals, and minicapsules can studies in animals in which intestinal chylomicron
be used for administration to small animals (Li and formation and lymphatic transport is blocked via
Zhao, 2007; Shah and Agnihotri, 2011). Intravenous preadministration of small molecular inhibitors of
formulations may be administered via i.v. bolus or chylomicron assembly (Dahan and Hoffman, 2005),
infusion into a catheter or cannula. assessment of drug secretion within lipoproteins on
The importance of formulation digestion by gastric incubation with caco-2 cells (Seeballuck et al., 2003,
and pancreatic enzymes in the performance of LBF 2004; Karpf et al., 2006), in vitro assessment of drug
and the potential role of bile salt solubilization in the affinity for collected intestinal triglyceride-rich li-
prevention of drug precipitation dictate that differ- poproteins (Gershkovich and Hoffman, 2005; Trevaskis
ences in physiology associated with bile salt release et al., 2010b), and in silico models that rely on simple
and digestion profile may affect product performance. molecular descriptors such as log P, long-chain TG
Notable differences are evident, for example, in solubility, and molecular weight (Holm and Hoest,
biliary secretory patterns in rats (where bile flow is 2004).
continuous) and in dogs or humans, where bile is
stored in the gallbladder and released as a bolus in D. Summary
response to lipid sensing in the GI tract. A consensus Lipids may be formulated into a range of de-
view as to whether these differences lead to higher livery systems for oral or parenteral administration
absorption in dogs or rats, however, has yet to emerge, (solutions, suspensions, emulsions, microemulsions,
although apocryphal evidence suggests that for poorly nanoemulsions, micellar solutions, liposomes, lipid
water-soluble drugs, absorption may be higher in dogs nanoparticles, and emulsion preconcentrates). LBF
than in rats. provide a relatively facile means of generating liquid
A second significant variable in the in vivo evalua- formulations that can be gavaged at varying doses
tion of drug absorption from LBF is the potential for during preclinical development and that provide for
intestinal lymphatic drug transport. For highly lipo- significant increases in exposure for many poorly
philic drugs, assessment of the extent of lymphatic water-soluble drugs. In addition, LBF have the
transport may be important, as small changes in advantage that essentially the same fill material can
formulation components and the presence of food may be subsequently filled into soft or hard gelatin capsules
change the extent of lymphatic transport and in doing to facilitate clinical development. LBF can also be
so impact drug exposure, activity, and toxicity via transformed into solid dosage forms directly via the use
changes to drug clearance and distribution (Porter and of high melting point lipids or polymers or via
Charman, 2001c,d; O’Driscoll, 2002; Porter et al., 2007; adsorption onto high-surface-area carriers.
Trevaskis et al., 2008; Caliph et al., 2012). To this Oral LBF commonly include combinations of glycer-
point, intestinal lymphatic drug transport has not been ide lipids, lipophilic or hydrophilic surfactants, and
measured directly in humans as the surgical cannula- cosolvents, resulting in formulations that spontane-
tion of the mesenteric lymph duct is highly invasive ously emulsify on contact with GI fluids (so called self-
(and typically nonreversible). However, several animal emulsifying drug delivery systems, or SEDDS). In the
models have been described for the assessment of preferred embodiment of the technology, drug is
lymphatic drug transport after oral administration. present in a dissolved form within a nonaqueous
These models involve the collection of lymph flowing solution in the fill material. Under these circum-
from the intestine via insertion of a cannula into either stances, traditional dissolution is circumvented and
the mesenteric or the thoracic lymph duct and sub- drug absorption proceeds via dispersion of the dose
sequent assessment of drug uptake into the lymph form in the GI content and integration into endogenous
after oral or intestinal administration to conscious or lipid digestion and processing pathways. Alternatively,
anesthetized animals. Studies have been conducted in where drug dose requirements exceed the solubility
rats (Edwards et al., 2001), dogs (Khoo et al., 2001, limit in the formulation, lipid suspensions may be
2003; Lespine et al., 2006), pigs (White et al., 1991), used, although this provides additional complexity
sheep (Onizuka et al., 1997; Segrave et al., 2004), and with respect to content uniformity and reinstalls the
rabbits (Bocci et al., 1986). The advantages and need to overcome crystal lattice energy limitations to
disadvantages and variations in conduct of lymphatic solubility.
drug transport studies in animals are summarized by LBF have the potential to support oral bioavailabil-
Edwards et al. (2001) and have been discussed ity via several mechanisms—including enhancements
472 Williams et al.

in drug dissolution and solubility (via stimulation of water-soluble drugs, but these strategies have yet to
bile salt release and incorporation of drug and lipid progress to broad clinical application. These are
digestion products into intestinal mixed micelles with described in brief in the sections below.
enhanced solubilization capacity), promotion of in-
testinal lymphatic drug transport (and attendant A. Drug Adsorption to Microporous Adsorbents
increases in oral bioavailability via decreases in first The use of microporous adsorbents has found in-
pass metabolism), and direct inhibition of intestinal creasing recent interest as a mechanism to facilitate
drug efflux or metabolism. Recent data further suggest the oral delivery of amorphous drug. The earliest
the potential for LBF to generate and sustain super- reports of an improved dissolution rate of poorly
saturation in the GI tract leading to decreased soluble drugs using adsorbents such as fumed silica
precipation, increased thermodynamic activity, and dioxide emerged in the 1970s (Monkhouse and Lach,
increased absorption (Gao et al., 2004; Anby et al., 1972a,b). More recent studies have extended these
2012; Williams et al., 2012b). early concepts to focus on drug adsorption to materials
LBF are also used to facilitate paren- with increasingly high surface areas, such as micropo-
teral administration of highly lipophilic drugs where rous silica, and have shown significant in vitro utility
traditional formulation options (e.g., pH manipulation, (Salonen et al., 2005; Ambrogi et al., 2007; Pan et al.,
cosolvents, cyclodextrins) are unable to provide for 2008; Van Speybroeck et al., 2009; Wang et al., 2010a;
sufficient solubilization. In the preclinical setting, this Vialpando et al., 2011; Ambrogi et al., 2012). Enhance-
is usually achieved by incorporation of drug into ments in oral bioavailability have been described after
commercial (preformed) parenteral lipid emulsions. oral administration of adsorbed microporous systems
For clinical applications, drug is dissolved in the lipid compared with unmodified crystalline material with
components that comprise the dispersed phase of the model compounds, including itraconazole (Mellaerts
formulation and emulsions are subsequently generated et al., 2008; Van Speybroeck et al., 2010b), fenofibrate
de novo. Parenteral lipid emulsions may, however, (Van Speybroeck et al., 2010a; Jia et al., 2011),
impact drug clearance and biodistribution profiles, and glibenclamide (Van Speybroeck et al., 2011), indometh-
this requires careful consideration when conducting acin (Wang et al., 2010a), and K-832 (Miura et al.,
fundamental pharmacokinetic studies. 2010). Substantial increases in bioavailability have
The clinical and commercial utility of LBF is evi- been described, and in the latter example, a ;5-fold
denced by the availability of several marketed prod- enhancement in the oral bioavailability of the de-
ucts. Increasing application in preclinical absorption, velopment candidate K-832 (under development for
disposition, metabolism, and elimination studies is immune, inflammatory, and ischemic diseases) was
also apparent, to support improved exposure for poorly evident after oral administration of the drug adsorbed
water-soluble compounds, regardless of whether a LBF to a porous silica substrate compared with the un-
is envisaged as the final product. Increasing experience modified crystal form (Miura et al., 2010).
in preclinical testing will likely drive further confi- The ability of silica carriers to increase drug dis-
dence in the application of formulations of this type, solution (and therefore oral bioavailability) has been
especially during toxicity testing, where very high suggested to reflect their capacity to stabilize drug in
exposure is usually required. Realization that LBF are the more rapidly dissolving amorphous form (Mellaerts
highly integrated into lipid digestion/absorption path- et al., 2007; Prestidge et al., 2007; Van Speybroeck
ways after oral administration has also resulted in et al., 2009). Silica carrier systems are therefore anal-
a shift in formulation design criteria such that ogous to solid dispersion technologies (see Section X)
attention is now focused more on formulation proper- that derive benefit via delivery of drug in the
ties (in particular, solubilization capacity) after di- amorphous form. For this reason, silica carrier systems
gestion and interaction with bile salt micelles rather have been described as “surface solid dispersions”
than the intrinsic properties of the formulation on the (Kerc et al., 1998). The mechanisms of amorphous
bench (which may be lost on processing in vivo). drug stabilization across the two delivery technologies,
Refinements to in vitro dispersion and digestion however, are different: solid dispersion formulations
testing protocols have greatly assisted this endeavor typically contain drug molecularly dispersed or suspended
by providing relatively simple tools with which to throughout an inert carrier (usually a high-molecular-
assess likely changes in performance in vivo. weight polymer), which slows drug recrystallization
via an increase in viscosity and a decrease in drug
mobility within the matrix. In contrast, amorphous
XII. Emerging Strategies for Improving the
drug adsorbed to the surfaces of a porous carrier is
Aqueous Solubility of Poorly Water-
stabilized via spatial confinement to small pores or
Soluble Drugs
capillaries that restrict drug nucleation and crystal
A number of strategies have emerged in recent years growth as a result of the overall small size of these
with the potential to enhance the delivery of poorly pores (Mellaerts et al., 2007; Wang et al., 2010a; Bras
Strategies for Low Drug Solubility 473

et al., 2011). Drug crystallization tendency is also rapid drug release from compressed silica formulations
decreased by the formation of drug-carrier interactions (Vialpando et al., 2011).
(Gupta et al., 2003; Van Speybroeck et al., 2009; Bras The physical stability challenges associated with
et al., 2011), although the formation of such inter- high-energy formulations must also be considered
actions is dependent on drug positioning within the when using porous silica carriers containing amor-
pore/capillary (Bras et al., 2011) and the chemical phous drug. In particular, the potential for recrystal-
surface of the silica carrier (Madieh et al., 2007; Qian lization of the stable crystal form over time, during
and Bogner, 2012). processing, and on exposure to moisture and after
The most effective carriers for amorphous drug dissolution. These physical stability considerations are
delivery provide an extensive network of nanosized or analogous to those described for solid dispersions (see
microsized capillaries or pores, resulting in high sur- Section X).
face areas (usually .500 m2/g) and high pore volumes
(usually around 1 cm3/g); properties that maximize B. Solid Lipid Nanoparticles
drug loading in the amorphous form (Vialpando et al., Solid lipid nanoparticles (SLNs) consist of a lipidic
2011; Qian and Bogner, 2012). Drug is commonly core that is solid at both room temperature and
loaded onto the carrier via solvent impregnation, where physiologic temperatures, and a surfactant-stabilized
a known volume of a concentrated solution of drug in outer surface (Mehnert and Mader, 2001; Bummer,
organic solvent is added to the carrier before mixing and 2004; Muller et al., 2011b). The concept of using solid
solvent evaporation at elevated temperatures. This particulate lipid vehicles for drug delivery emerged
provides a dry, free-flowing powder of known drug more than 20 years ago (Eldem et al., 1991), and a
loading (Mellaerts et al., 2007; Van Speybroeck et al., significant level of interest has subsequently focused
2009). Alternatively, drug-carrier physical mixtures on the use of SLN formulations for 1) improved oral
may be dry milled (Gupta et al., 2003; Bahl and Bogner, bioavailability of poorly water-soluble drugs (Hu et al.,
2006) or drug loaded via vapor adsorption (Qian and 2004c; Luo et al., 2006; Muller et al., 2006), 2)
Bogner, 2011; Qian et al., 2011). The advantages and parenteral administration of poorly water-soluble
disadvantages of alternative drug-loading approaches drugs (Wissing et al., 2004; Manjunath and Venkates-
have been described by Qian and Bogner (2012). warlu, 2005; Reddy et al., 2005), 3) enhanced topical
Silica-based carriers constitute the most widely in- drug delivery, 4) controlled drug release, and 5) targeted
vestigated porous materials in drug delivery. Commer- drug delivery. Descriptions of SLN formulations for
cial examples include magnesium aluminometasilicate dermal application or sustained and targeted drug re-
(Neuselin; Fuji Chemical Industry Co., Ltd., Tokyo, lease are beyond the scope of the present review, but
Japan), calcium silicate (Florite; Tokuyama Corpora- they have been extensively reviewed (zur Mühlen et al.,
tion, Tokyo, Japan), and ordered mesoporous silicon 1998; Boyd, 2008; Pardeike et al., 2009).
dioxide (Solustia; Formac Pharmaceuticals, Heverlee, SLNs combine the advantages of colloidal delivery
Belgium). Ordered mesoporous silicas (OMS), including systems such as liposomes, lipid emulsions, and poly-
SBA-15 and MCM-41, show a collection of unique meric nanoparticles (Müller et al., 2000; Wissing et al.,
properties that are especially attractive to drug deliv- 2004), with the potential for facile large-scale pro-
ery, such as very high specific surface areas (up to 1500 duction and a favorable tolerability profile for most
m2/g), large pore volumes (up to 1.5 g/cm3), a tuneable core lipid components (Müller et al., 2000; Bummer,
pore diameter (for variable drug release rates), and 2004). The drug-loading capacity of SLNs for lipid-
functionalized surfaces (Van Speybroeck et al., 2009; soluble drugs may be high (Schwarz and Mehnert,
Popovici et al., 2011). Polymeric precipitation inhibitors 1999); however, for less lipid-soluble compounds, low
(HPMC and HPMCAS) may also be used in combination drug loading may limit applicability (Mehnert and
with OMS to slow the rate of recrystallization of the Mader, 2001). Second-generation SLNs, usually termed
supersaturated drug solutions that form on dissolution nanostructured lipid carriers (NLCs), have subse-
of the amorphous drug load (Van Speybroeck et al., quently been developed. NLCs incorporate blends of
2010b). solid and liquid lipids (but retain solid properties at
Although there is significant interest in the use of room temperature) that increase the drug-loading capa-
microporous materials as carriers for the delivery of city of the lipid core (Muller et al., 2002). For example,
poorly water-soluble drugs, in particular, to stabilize the addition of liquid oils to an SLN formulation of
amorphous drug content, the feasibility of manufactur- retinol increases the maximum drug loading from 1%
ing silica-based formulations at larger production w/w (using only glyceryl behenate) to 5% (using a com-
scales is not yet known (Qian and Bogner, 2012). The bination of glyceryl behenate and Miglyol 812) (Jenning
processability of formulations that contain high (usu- and Gohla, 2001).
ally .80% w/w) quantities of silica-based materials are SLN formulations contain drug molecularly dis-
also not fully described, and recent work suggests that persed within the lipidic vehicle. Drug delivery via
diluents and disintegrants may be necessary to ensure SLN formulations, therefore, circumvents the potentially
474 Williams et al.

slow rate of dissolution of poorly water-soluble drugs should also be considered. Similar to drug nanoparticle
from crystalline solids (Muller et al., 2006). Cyclospor- formulations (see section Section IX), stabilizers are
ine (Muller et al., 2006), nitrendipine (Manjunath and required to facilitate the production of nanosized
Venkateswarlu, 2006), puerarin (Luo et al., 2011), material and to maintain the long-term physical
quercetin (Li et al., 2009a), and lovastatin (Suresh stability of SLN formulations (Mehnert and Mader,
et al., 2007) all show increases in oral bioavailability 2001; Bummer, 2004). Combinations of different stabil-
when delivered in SLN formulations compared with izers are often used to stabilize SLN formulations ag-
the administration of more traditional suspension and ainst potential agglomeration. Common emulsifiers include
nanosuspension formulations. In some cases, SLN polysorbate surfactants (e.g., Tweens), poloxamers (e.g.,
formulations have also been directly compared with Pluronics), lecithin, and various bile salts.
alternate solubilization techniques such as cosolvent- Usually, SLNs are produced via high-pressure homo-
surfactant solutions (Hu et al., 2004c; Luo et al., 2006) genization techniques (Mehnert and Mader, 2001).
and lipid emulsions (Hu et al., 2004c) and shown to These are summarized in Section IX. During “hot”
provide for superior bioavailability after oral adminis- homogenization, drug/lipid melts are mixed with a hot
tration. SLN formulations, like traditional lipid-based aqueous surfactant solution and then passed through
drug delivery formulations, are expected to be digested a homogenizer while the temperature is maintained
in the small intestine (Muller et al., 2011b). It seems approximately 5°C higher than the melting tempera-
likely, therefore, that a better understanding of the ture of the lipid component(s). The resultant nano-
factors that influence in vivo performance of SLNs emulsion is cooled to room temperature, causing the
may be obtained via the application of in vitro lipid lipid matrix to crystallize and form solid nanoparticles.
digestion models (similar to those described in detail in “Cold homogenization” techniques have been devel-
the lipid formulation section of this review; see Section oped to minimize potential temperature-induced drug
XI) to evaluate better the impact of digestion on drug degradation and drug partitioning from the oil phase
solubilization. into surrounding aqueous phase (Mehnert and Mader,
In addition, SLN formulations may be used to 2001). In this method, the drug-lipid melts are rapidly
facilitate parenteral delivery of poorly water-soluble cooled with liquid nitrogen, and particles are milled to
drugs (Wissing et al., 2004; Reddy et al., 2005; Muller form coarse lipid microparticles. The microparticles are
et al., 2011b) and in a manner analogous to nanocrystal subsequently dispersed in cold surfactant solution and
suspensions (see Section IX), aqueous dispersions may subjected to high-pressure homogenization to reduce
be lyophilized for convenient transport and storage and particle size (Müller et al., 2000). SLNs can be
later reconstituted before use (Zimmermann et al., manufactured in large scale using high-pressure ho-
2000). Recent interest in the parenteral utility of SLN mogenization (Muller et al., 2011b).
formulations has focused largely on use as a controlled- In addition, SLNs may be isolated by solvent
release platform (Cavalli et al., 1999; Mehnert and evaporation; lipids are dissolved in water-immiscible
Mader, 2001; Jia et al., 2010). In this application, drug organic solvents, mixed with an aqueous surfactant
partitioning from the oil phase into the surrounding solution, and the solvent evaporated (Sjostrom and
aqueous environments is expected to be hindered com- Bergenstahl, 1992); and by combined ultrasonication-
pared with liquid nanoemulsion oil droplets (of equal solvent evaporation techniques. Ultrasonication-solvent
diameter) (zur Mühlen et al., 1998; Mehnert and techniques mix drug, lipid, and organic solvent at high
Mader, 2001) by virtue of the lower mobility of the temperature before solvent evaporation and drop-wise
drug within the lipid component of the formulation. addition of an aqueous surfactant solution to the warm
SLN formulations also provide opportunities for tar- lipid phase. Ultrasonication of the subsequent coarse
geted therapies; however, these lie beyond the scope of emulsions and dispersion in water results in the for-
this review [see Muller et al. (2011b) and Joshi and mulation of nanoparticles (Luo et al., 2006). These
Müller (2009) for details]. techniques have been reviewed elsewhere (Mehnert and
Lipidic excipients widely used in SLN formulations Mader, 2001; Bummer, 2004).
are triglycerides (e.g., tricaprin and tristearin), mix- Since delivery of drug in a noncrystalline, molecu-
tures of partially digested glycerides (e.g., glyceryl larly dispersed form is critical to the utility of SLNs (at
mono-oleate or dioleate or stearate), fatty acids (e.g., least in applications in the area of low drug solubility),
stearic acid), cholesterol, and typical pharmaceutical evidence of drug crystallinity in SLNs after manufac-
waxes (e.g., cetyl palmitate). Since the core excipients ture or prolonged periods of storage is routinely
are usually GRAS-status lipids, SLN formulations offer assessed by DSC.
toxicological advantages over solubilization techniques
that require more complex components, particularly in
XIII. Conclusions
parenteral drug delivery. However, whereas the com-
ponents of the lipid core may be GRAS, the pharma- Although major progress has been made in the
cological side effects from the surfactant stabilizers application of “developability” criteria to support
Strategies for Low Drug Solubility 475

progression of a drug candidate, continuing efforts to Finally, a distinction should be made between
identify potent drug leads and a desire to explore hydrophobic drugs (typically high log P and low lipid
nontraditional drug targets dictate that low water solubility) and truly lipophilic drugs (high log P and
solubility is an ever-present challenge in drug dis- high lipid solubility). This difference may also be
covery and development. In response, an ever in- couched in terms of solvation and solid-state proper-
creasing arsenal of approaches has been developed to ties, where poorly water-soluble hydrophobic drugs
address the challenge of low water solubility, along (colloquially referred to as “brick dust”) show both
with continued improvement in the depth of un- solvation and solid-state property limited solubility
derstanding of the principles that underpin these and poorly water-soluble lipophilic drugs (“grease
technologies. balls”) are typically solvation-limited in water, but
The determinants of low aqueous solubility are poor the lack of solid-state limitations and good solvation in
solvation (i.e., unfavorable solute-solvent interactions nonaqueous vehicles promote lipid solubility. Lipophilic
in solution) and high crystal lattice energy (i.e., strong compounds are more readily delivered using formu-
solute-solute interactions in the solid state). Approaches lations typically grouped as lipid-based formulations,
to address apparent solubility therefore seek to enhance which may contain a wide range of nonaqueous and
solute-solvent interactions or to reduce solute-solute usually liquid vehicles, including glyceride lipids, sur-
interactions in the solid state. Considerable progress factants, and cosolvents. When delivered as a nonaque-
has been made in the application of a wide range of ous solution filled into gelatin capsules, LBF have the
technologies to address these issues. Increasingly flex- advantage of circumventing traditional dissolution pro-
ible and robust formulation (solid dispersions, lipid- cesses and integrating into endogenous (and highly
based formulations, nanocrystals, cyclodextrins) and efficient) lipid absorption pathways. However, LBF
chemical form (salts, co-crystals) approaches are appar- must be carefully assessed to understand changes in
ent, most of which have been applied to clinically and properties as the formulation is processed by digestion
commercially successful products. in the GI tract and are most effective for drugs where
Progress has also been made in understanding lipid solubility is high.
fundamental solute-solvent interactions, and de novo In summary, it seems likely that the challenge of
prediction of indicators of lipophilicity (log P) and low water solubility in drug hits, leads, and de-
ionization (pKa) are now widespread. An ability to velopment candidates will continue for the foresee-
predict crystal lattice energy from first principles, able future and, therefore, that strategies to address
however, is still elusive and continues to hamper pre- these issues will remain a critical determinant of
dictions of utility for approaches predicated on changes product success (or failure). The techniques overviewed
to solid-state properties. These include a priori pre- here provide a framework from which to build formula-
dictions of Ksp (solubility product constant) for differ- tion approaches at all stages of development, and such
ent salt forms or cocrystals and the likely impact of approaches have been applied in tried and tested drug
solid-state properties on solid cyclodextrin complexes products. Nonetheless, challenges remain, and many
or solid dispersion formation. Under these circum- highly potent and highly selective molecules still have
stances, high-throughput screening approaches are not progressed to development because of their low
used to select the most advantageous approach rather solubility. Better tools are required to ensure re-
than the application of bottom-up rational design. producible exposure, especially during toxicity testing,
The inherent advantage of high-energy solids (e.g., where very large doses are required. To this end,
the amorphous form) in enhancing solubility, at least considerable effort within the pharmaceutical sciences
over short periods, has long held appeal. Practical remains focused on better understanding the relation-
application of these approaches, however, has been ships between molecular structure and solubility and
limited by concerns over prediction of the kinetics of on the development of novel approaches to overcome
recrystallization and, therefore, the approaches that the barriers to solubility and to facilitate clinically
might be taken to slow recrystallization over phar- useful exposure.
maceutically relevant time frames. Advances in un-
derstanding of the drivers of recrystallization and Acknowledgments
correlation with readily measurable thermal properties The authors thank Professor Naίr Rodrίguez-Hornedo for insight
(notably Tg) have significantly increased confidence and clarification of several issues associated with the use of
pharmaceutical cocrystals. The authors also thank the reviewers of
with these approaches, and increasing numbers of
this manuscript for their comments. The authors received no
marketed products provide evidence that the formula- financial support for the research and/or authorship of this review
tion of drugs in the amorphous state is a practical article.
option. Concerns remain, however, over the importance
of microstructure, how best to fingerprint, and the Authorship Contributions
difficulties in predicting stability in systems that do Wrote or contributed to the writing of the manuscript: Williams,
not follow Arrhenius kinetics. Trevaskis, S. A. Charman, Shanker, W. N. Charman, Pouton, Porter.
476 Williams et al.

References of the impact of supersaturation stabilization on the in vitro and in vivo perfor-
Aakeröy CB, Fasulo ME, and Desper J (2007) Cocrystal or salt: does it really matter? mance of self-emulsifying drug delivery systems. Mol Pharm 9:2063–2079.
Anderberg EK, Bisrat M, and Nystrom C (1988) Physicochemical aspects of drug
Mol Pharm 4:317–322.
Aakeröy CB and Salmon DJ (2005) Building co-crystals with molecular sense and release. 7. The effect of surfactant concentration and drug particle-size on solubility
supramolecular sensibility. Cryst Eng Comm 7:439–448. and dissolution rate of felodipine, a sparingly soluble drug. Int J Pharm 47:67–77.
Aakeröy CB and Schultheiss N (2007) Assembly of molecular solids via non-covalent Anderberg EK, Nyström C, and Artursson P (1992) Epithelial transport of drugs in
interactions, in Making Crystals by Design: Methods, Techniques and Applications cell culture. VII: Effects of pharmaceutical surfactant excipients and bile acids on
(Braga D and Grepioni F eds) pp 209–240, Wiley-VCH, Weinheim, Germany. transepithelial permeability in monolayers of human intestinal epithelial (Caco-2)
Abdalla A, Klein S, and Mäder K (2008) A new self-emulsifying drug delivery system cells. J Pharm Sci 81:879–887.
(SEDDS) for poorly soluble drugs: characterization, dissolution, in vitro digestion Andersen FA (1994) Final report on the safety assessment of propylene-glycol and
and incorporation into solid pellets. Eur J Pharm Sci 35:457–464. polypropylene glycols. J Am Coll Toxicol 13:437–491.
Abdalla A and Mäder K (2009) ESR studies on the influence of physiological disso- Anderson BD and Conradi RA (1985) Predictive relationships in the water solubility
lution and digestion media on the lipid phase characteristics of SEDDS and of salts of a nonsteroidal anti-inflammatory drug. J Pharm Sci 74:815–820.
SEDDS pellets. Int J Pharm 367:29–36. Andronis V and Zografi G (1997) Molecular mobility of supercooled amorphous in-
Abu-Diak OA, Jones DS, and Andrews GP (2011) An investigation into the dissolu- domethacin, determined by dynamic mechanical analysis. Pharm Res 14:410–414.
tion properties of celecoxib melt extrudates: understanding the role of polymer type Andronis V and Zografi G (1998) The molecular mobility of supercooled amorphous
and concentration in stabilizing supersaturated drug concentrations. Mol Pharm 8: indomethacin as a function of temperature and relative humidity. Pharm Res 15:
1362–1371. 835–842.
Adam A, Schrimpl L, and Schmidt PC (2000) Factors influencing capping and Angell CA (1995a) Formation of glasses from liquids and biopolymers. Science 267:
cracking of mefenamic acid tablets. Drug Dev Ind Pharm 26:489–497. 1924–1935.
Adams ML, Lavasanifar A, and Kwon GS (2003) Amphiphilic block copolymers for Angell CA (1995b) The old problems of glass and the glass transition, and the many
drug delivery. J Pharm Sci 92:1343–1355. new twists. Proc Natl Acad Sci USA 92:6675–6682.
Adrjanowicz K, Kaminski K, Paluch M, Wlodarczyk P, Grzybowska K, Wojnarowska Z, Angell CA (2011) Heat capacity and entropy functions in strong and fragile glass-
Hawelek L, Sawicki W, Lepek P, and Lunio R (2010a) Dielectric relaxation formers, relative to those of disordering crystalline materials, in Hot Topics in
studies and dissolution behavior of amorphous verapamil hydrochloride. J Pharm Sci Thermal Analysis and Calorimetry #8: Glassy, Amorphous and Nano-Crystalline
99:828–839. Materials: Thermal Physics, Analysis, Structure and Properties (Sestak J, Mares
Adrjanowicz K, Paluch M, and Ngai KL (2010b) Determining the structural re- JJ, and Hubik P eds) pp 21–40, Springer, New York.
laxation times deep in the glassy state of the pharmaceutical telmisartan. J Phys Angell CA and Tucker JC (1980) Heat-capacity changes in glass-forming aqueous-
Condens Matter 22:125902. solutions and the glass-transition in vitreous water. J Phys Chem 84:268–272.
Agarwal V, Siddiqui A, Ali H, and Nazzal S (2009) Dissolution and powder flow Angell CA, Yue YZ, Wang LM, Copley JRD, Borick S, and Mossa S (2003) Potential
characterization of solid self-emulsified drug delivery system (SEDDS). Int J energy, relaxation, vibrational dynamics and the boson peak, of hyperquenched
Pharm 366:44–52. glasses. J Phys Condens Matter 15:S1051–S1068.
Agellon LB, Toth MJ, and Thomson ABR (2002) Intracellular lipid binding proteins of Anonymous (2005) 125 Questions: What we don’t know? Science 309:75.
the small intestine. Mol Cell Biochem 239:79–82. Antlsperger G and Schmid G (1996) Toxicological comparison of cyclodextrins, in
Agharkar S, Lindenbaum S, and Higuchi T (1976) Enhancement of solubility of drug Proceedings of the 8th International Symposium on Cyclodextrins (Szejtli J
salts by hydrophilic counterions: properties of organic salts of an antimalarial drug. and Szente L eds) Kluwer Academic Publishers; 1996 March 31 and April 2;
J Pharm Sci 65:747–749. Budapest, Hungary.
Agoram B, Sagawa K, Shanker RM, and Singh SK (2010) Biopharmaceutics of NCEs Aramwit P, Kerdcharoen T, and Qi H (2006) In vitro plasma compatibility study of
and NBEs, in Pharmaceutical Dosage Forms: Parenteral Medications (Nema S a nanosuspension formulation. PDA J Pharm Sci Technol 60:211–217.
and Ludwig JD eds) Informa Healthcare, New York. Arias MJ, Gines JM, Moyano JR, Perez-Martinez JI, and Rabasco AM (1995) In-
Agüeros M, Zabaleta V, Espuelas S, Campanero MA, and Irache JM (2010) Increased fluence of the preparation method of solid dispersions on their dissolution rate:
oral bioavailability of paclitaxel by its encapsulation through complex formation study of triamterene-D-mannitol system. Int J Pharm 123:25–31.
with cyclodextrins in poly(anhydride) nanoparticles. J Control Release 145:2–8. Arima H, Yunomae K, Hirayama F, and Uekama K (2001a) Contribution of P-
Aguiar AJ, Krc J Jr, Kinkel AW, and Samyn JC (1967) Effect of polymorphism on the glycoprotein to the enhancing effects of dimethyl-beta-cyclodextrin on oral bio-
absorption of chloramphenicol from chloramphenicol palmitate. J Pharm Sci 56: availability of tacrolimus. J Pharmacol Exp Ther 297:547–555.
847–853. Arima H, Yunomae K, Miyake K, Irie T, Hirayama F, and Uekama K (2001b)
Aguiar AJ and Zelmer JE (1969) Dissolution behavior of polymorphs of chloram- Comparative studies of the enhancing effects of cyclodextrins on the solubility and
phenicol palmitate and mefenamic acid. J Pharm Sci 58:983–987. oral bioavailability of tacrolimus in rats. J Pharm Sci 90:690–701.
Ahmed K, Li Y, McClements DJ, and Xiao H (2012) Nanoemulsion- and emulsion- Arima H, Yunomae K, Morikawa T, Hirayama F, and Uekama K (2004) Contribution
based delivery systems for curcumin: encapsulation and release properties. Food of cholesterol and phospholipids to inhibitory effect of dimethyl-beta-cyclodextrin
Chem 132:799–807. on efflux function of P-glycoprotein and multidrug resistance-associated protein 2
Ahmed MO (2001) Comparison of impact of the different hydrophilic carriers on the in vinblastine-resistant Caco-2 cell monolayers. Pharm Res 21:625–634.
properties of piperazine-containing drug. Eur J Pharm Biopharm 51:221–225. Arimori K and Uekama K (1987) Effects of beta- and gamma-cyclodextrins on the
Ainouz A, Authelin JR, Billot P, and Lieberman H (2009) Modeling and prediction of pharmacokinetic behavior of prednisolone after intravenous and intramuscular
cocrystal phase diagrams. Int J Pharm 374:82–89. administrations to rabbits. J Pharmacobiodyn 10:390–395.
Albers E and Müller BW (1992) Complexation of steroid hormones with cyclodextrin Armstrong J, Chowdhry B, Mitchell J, Beezer A, and Leharne S (1996) Effect of
derivatives: substituent effects of the guest molecule on solubility and stability in cosolvents and cosolutes upon aggregation transitions in aqueous solutions of the
aqueous solution. J Pharm Sci 81:756–761. poloxamer F87 (poloxamer P237): a high sensitivity differential scanning calo-
Alie J, Menegotto J, Cardon P, Duplaa H, Caron A, Lacabanne C, and Bauer M (2004) rimetry study. J Phys Chem 100:1738–1745.
Dielectric study of the molecular mobility and the isothermal crystallization ki- Arnarson T and Elworthy PH (1982) Effects of structural variations of non-ionic
netics of an amorphous pharmaceutical drug substance. J Pharm Sci 93:218–233. surfactants on micellar properties and solubilization: variation of oxyethylene
Allen FH (2002) The Cambridge Structural Database: a quarter of a million crystal content on properties of C22 monoethers. J Pharm Pharmacol 34:87–89.
structures and rising. Acta Crystallogr B 58:380–388. Arnesjö B and Grubb A (1971) The activation, purification and properties of rat
Allen LV, Levinson RS, and Martono DD (1978) Dissolution rates of hydrocortisone pancreatic juice phospholipase A 2. Acta Chem Scand 25:577–589.
and prednisone utilizing sugar solid dispersion systems in tablet form. J Pharm Sci Arora KK and Zaworotko MJ (2009) Pharmaceutical co-crystals: a new opportunity in
67:979–981. pharmaceutical science for a long-known but little-studied class of compounds, in
Al-Marzouqi AH, Shehatta I, Jobe B, and Dowaidar A (2006) Phase solubility and Polymorphism in Pharmaceutical Solids (Brittain HG ed) pp 282–317, Informa
inclusion complex of itraconazole with beta-cyclodextrin using supercritical carbon Healthcare, New York.
dioxide. J Pharm Sci 95:292–304. Aso Y, Yoshioka S, and Kojima S (2000) Relationship between the crystallization
Alonzo DE, Gao Y, Zhou DL, Mo HP, Zhang GGZ, and Taylor LS (2011) Dissolution rates of amorphous nifedipine, phenobarbital, and flopropione, and their molecular
and precipitation behavior of amorphous solid dispersions. J Pharm Sci 100: mobility as measured by their enthalpy relaxation and (1)H NMR relaxation times.
3316–3331. J Pharm Sci 89:408–416.
Alsenz J and Kansy M (2007) High throughput solubility measurement in drug Attwood D and Florence AT (1983) Surfactant Systems: Their Chemistry, Pharmacy
discovery and development. Adv Drug Deliv Rev 59:546–567. and Biology, Chapman and Hall, London.
Alvarez-Núñez FA and Yalkowsky SH (2000) Relationship between polysorbate 80 Augustijns P and Brewster ME (2012) Supersaturating drug delivery systems: fast is
solubilization descriptors and octanol-water partition coefficients of drugs. Int J not necessarily good enough. J Pharm Sci 101:7–9.
Pharm 200:217–222. Aungst BJ (2000) Intestinal permeation enhancers. J Pharm Sci 89:429–442.
Ambrogi V, Perioli L, Marmottini F, Giovagnoli S, Esposito M, and Rossi C (2007) Auzely-Velty R, Djedaini-Pilard F, Desert S, Perly B, and Zemb T (2000) Micellization
Improvement of dissolution rate of piroxicam by inclusion into MCM-41 mesoporous of hydrophobically modified cyclodextrins. 1. Micellar structure. Langmuir 16:
silicate. Eur J Pharm Sci 32:216–222. 3727–3734.
Ambrogi V, Perioli L, Pagano C, Latterini L, Marmottini F, Ricci M, and Rossi C Avis KE, Lachman L, and Lieberman HA (1986) Pharmaceutical Dosage Forms:
(2012) MCM-41 for furosemide dissolution improvement. Microporous Mesoporous Parenteral Medications, Marcel Dekker, New York.
Mater 147:343–349. Badawy SIF, Ghorab MM, and Adeyeye CM (1996) Characterization and bio-
Amidon GL, Lennernäs H, Shah VP, and Crison JR (1995) A theoretical basis for availability of danazol-hydroxypropyl beta-cyclodextrin coprecipitates. Int J Pharm
a biopharmaceutic drug classification: the correlation of in vitro drug product 128:45–54.
dissolution and in vivo bioavailability. Pharm Res 12:413–420. Badens E, Majerik V, Horváth G, Szokonya L, Bosc N, Teillaud E, and Charbit G
Amin K and Dannenfelser RM (2006) In vitro hemolysis: guidance for the pharma- (2009) Comparison of solid dispersions produced by supercritical antisolvent and
ceutical scientist. J Pharm Sci 95:1173–1176. spray-freezing technologies. Int J Pharm 377:25–34.
Anby MU, Williams HD, McIntosh M, Benameur H, Edwards GA, Pouton CW, Badrinarayanan P, Zheng W, Li Q, and Simon SL (2007) The glass transition tem-
and Porter CJH (2012) Lipid digestion as a trigger for supersaturation: evaluation perature versus the fictive temperature. J Non-Cryst Solids 353:2603–2612.
Strategies for Low Drug Solubility 477
Baert L, van ’t Klooster G, Dries W, Wouterss A, Basstanie E, Iterbeke K, Stappers F, Bettinetti G, Melani F, Mura P, Monnanni R, and Giordano F (1991) Carbon-13
Stevens P, Schueller L, and Van Remoortere P, et al. (2009) Development of a long- nuclear magnetic resonance study of naproxen interaction with cyclodextrins in
acting injectable formulation with nanoparticles of rilpivirine (TMC278) for HIV solution. J Pharm Sci 80:1162–1170.
treatment. Eur J Pharm Biopharm 72:502–508. Bevernage J, Forier T, Brouwers J, Tack J, Annaert P, and Augustijns P (2011)
Bahadur P, Li P, Almgren M, and Brown W (1992) Effect of potassium fluoride on the Excipient-mediated supersaturation stabilization in human intestinal fluids. Mol
micellar behavior of pluronic F-68 in aqueous-solution. Langmuir 8:1903–1907. Pharm 8:564–570.
Bahl D and Bogner RH (2006) Amorphization of indomethacin by co-grinding with Bhakay A, Merwade M, Bilgili E, and Dave RN (2011) Novel aspects of wet milling for
neusilin US2: amorphization kinetics, physical stability and mechanism. Pharm the production of microsuspensions and nanosuspensions of poorly water-soluble
Res 23:2317–2325. drugs. Drug Dev Ind Pharm 37:963–976.
Bahrami M and Ranjbarian S (2007) Production of micro- and nano-composite par- Bharatiya B, Guo C, Ma JH, Hassan PA, and Bahadur P (2007) Aggregation and
ticles by supercritical carbon dioxide. J Supercrit Fluids 40:263–283. clouding behavior of aqueous solution of EO-PO block copolymer in presence of n-
Baird JA and Taylor LS (2011) Evaluation and modeling of the eutectic composition alkanols. Eur Polym J 43:1883–1891.
of various drug-polyethylene glycol solid dispersions. Pharm Dev Technol 16: Bhat PA, Dar AA, and Rather GM (2008) Solubilization capabilities of some cationic,
201–211. anionic, and nonionic surfactants toward the poorly water-soluble antibiotic drug
Baird JA and Taylor LS (2012) Evaluation of amorphous solid dispersion properties erythromycin. J Chem Eng Data 53:1271–1277.
using thermal analysis techniques. Adv Drug Deliv Rev 64:396–421. Bhat S, Leikin-Gobbi D, Konikoff FM, and Maitra U (2006) Use of novel cationic bile
Baird JA, Van Eerdenbrugh B, and Taylor LS (2010) A classification system to assess salts in cholesterol crystallization and solubilization in vitro. Biochim Biophys Acta
the crystallization tendency of organic molecules from undercooled melts. J Pharm 1760:1489–1496.
Sci 99:3787–3806. Bhattachar SN, Deschenes LA, and Wesley JA (2006) Solubility: it’s not just for
Bak A, Gore A, Yanez E, Stanton M, Tufekcic S, Syed R, Akrami A, Rose M, physical chemists. Drug Discov Today 11:1012–1018.
Surapaneni S, and Bostick T, et al.. (2008) The co-crystal approach to improve the Bhattacharya S and Suryanarayanan R (2009) Local mobility in amorphous
exposure of a water-insoluble compound: AMG 517 sorbic acid co-crystal charac- pharmaceuticals—characterization and implications on stability. J Pharm Sci 98:
terization and pharmacokinetics. J Pharm Sci 97:3942–3956. 2935–2953.
Balbach S and Korn C (2004) Pharmaceutical evaluation of early development can- Bhugra C and Pikal MJ (2008) Role of thermodynamic, molecular, and kinetic factors
didates “the 100 mg-approach.” Int J Pharm 275:1–12. in crystallization from the amorphous state. J Pharm Sci 97:1329–1349.
Banerjee R, Bhatt PM, Ravindra NV, and Desiraju GR (2005) Saccharin salts of Biradha K, Su CY, and Vittal JJ (2011) Recent developments in crystal engineering.
active pharmaceutical ingredients, their crystal structures, and increased water Cryst Growth Des 11:875–886.
solubilities. Cryst Growth Des 5:2299–2309. Bisrat M and Nystrom C (1988) Physicochemical aspects of drug release. 8. The
Banerjee SK, Jagannath C, Hunter RL, and Dasgupta A (2000) Bioavailability of relation between particle-size and surface specific dissolution rate in agitated
tobramycin after oral delivery in FVB mice using CRL-1605 copolymer, an in- suspensions. Int J Pharm 47:223–231.
hibitor of P-glycoprotein. Life Sci 67:2011–2016. Bittner B, González RCB, Isel H, and Flament C (2003) Impact of Solutol HS 15 on
Bansal SS, Kaushal AM, and Bansal AK (2007) Molecular and thermodynamic the pharmacokinetic behavior of midazolam upon intravenous administration to
aspects of solubility advantage from solid dispersions. Mol Pharm 4:794–802. male Wistar rats. Eur J Pharm Biopharm 56:143–146.
Barry BW and El Eini DID (1976) Solubilization of hydrocortisone, dexamethasone, Bittner B, Guenzi A, Fullhardt P, Zuercher G, González RC, and Mountfield RJ
testosterone and progesterone by long-chain polyoxyethylene surfactants. J Pharm (2002) Improvement of the bioavailability of colchicine in rats by co-administration
Pharmacol 28:210–218. of D-alpha-tocopherol polyethylene glycol 1000 succinate and a polyethoxylated
Barta CA, Sachs-Barrable K, Feng F, and Wasan KM (2008) Effects of mono- derivative of 12-hydroxy-stearic acid. Arzneimittelforschung 52:684–688.
glycerides on P-glycoprotein: modulation of the activity and expression in Caco-2 Bittner B and Mountfield RJ (2002) Intravenous administration of poorly soluble new
cell monolayers. Mol Pharm 5:863–875. drug entities in early drug discovery: the potential impact of formulation on
Bartolini A, Renzi G, Galli A, Aiello PM, and Bartolini R (1981) Eseroline: a new pharmacokinetic parameters. Curr Opin Drug Discov Devel 5:59–71.
antinociceptive agent derived from physostigmine with opiate receptor agonist Bitz C and Doelker E (1996) Influence of the preparation method on residual solvents
properties: experimental in vivo and in vitro studies on cats and rodents. Neurosci in biodegradable microspheres. Int J Pharm 131:171–181.
Lett 25:179–183. Black DD (2007) Development and physiological regulation of intestinal lipid absorp-
Bartolomei M, Bertocchi P, Antoniella E, and Rodomonte A (2006) Physico-chemical tion. I. Development of intestinal lipid absorption: cellular events in chylomicron
characterisation and intrinsic dissolution studies of a new hydrate form of diclofenac assembly and secretion. Am J Physiol Gastrointest Liver Physiol 293:G519–G524.
sodium: comparison with anhydrous form. J Pharm Biomed Anal 40:1105–1113. Black SN, Collier EA, Davey RJ, and Roberts RJ (2007) Structure, solubility,
Bastin RJ, Bowker MJ, and Slater BJ (2000) Salt selection and optimisation proce- screening, and synthesis of molecular salts. J Pharm Sci 96:1053–1068.
dures for pharmaceutical new chemical entities. Org Process Res Dev 4:427–435. Blagden N, de Matas M, Gavan PT, and York P (2007) Crystal engineering of active
Batrakova EV, Han HY, Alakhov VYu, Miller DW, and Kabanov AV (1998) Effects of pharmaceutical ingredients to improve solubility and dissolution rates. Adv Drug
pluronic block copolymers on drug absorption in Caco-2 cell monolayers. Pharm Deliv Rev 59:617–630.
Res 15:850–855. Bloedow DC and Hayton WL (1976) Effects of lipids on bioavailability of sulfisoxazole
Batrakova EV, Li S, Miller DW, and Kabanov AV (1999) Pluronic P85 increases acetyl, dicumarol, and griseofulvin in rats. J Pharm Sci 65:328–334.
permeability of a broad spectrum of drugs in polarized BBMEC and Caco-2 cell Bocci V, Muscettola M, Grasso G, Magyar Z, Naldini A, and Szabo G (1986) The
monolayers. Pharm Res 16:1366–1372. lymphatic route. 1. Albumin and hyaluronidase modify the normal distribution of
Bauer J, Spanton S, Henry R, Quick J, Dziki W, Porter W, and Morris J (2001) interferon in lymph and plasma. Experientia 42:432–433.
Ritonavir: an extraordinary example of conformational polymorphism. Pharm Res Bogardus JB and Blackwood RK Jr (1979) Solubility of doxycycline in aqueous so-
18:859–866. lution. J Pharm Sci 68:188–194.
Bearnot HR, Glickman RM, Weinberg L, Green PH, and Tall AR (1982) Effect of Bogman K, Erne-Brand F, Alsenz J, and Drewe J (2003) The role of surfactants in the
biliary diversion on rat mesenteric lymph apolipoprotein-I and high density lipo- reversal of active transport mediated by multidrug resistance proteins. J Pharm
protein. J Clin Invest 69:210–217. Sci 92:1250–1261.
Becher P (1967) Micelle formation in aqueous and nonaqueous solutions, in Nonionic Bogman K, Zysset Y, Degen L, Hopfgartner G, Gutmann H, Alsenz J, and Drewe J
Surfactants (Schick JM ed) pp 478–515, Marcel Dekker, New York. (2005) P-glycoprotein and surfactants: effect on intestinal talinolol absorption. Clin
Behrens D, Fricker R, Bodoky A, Drewe J, Harder F, and Heberer M (1996) Com- Pharmacol Ther 77:24–32.
parison of cyclosporin A absorption from LCT and MCT solutions following intra- Bom A, Bradley M, Cameron K, Clark JK, van Egmond J, Feilden H, MacLean EJ,
jejunal administration in conscious dogs. J Pharm Sci 85:666–668. Muir AW, Palin R, and Rees DC, et al. (2002) A novel concept of reversing neu-
Bekers O, Uijtendaal EV, Beijnen JH, Bult A, and Underberg WJM (1991) Cyclo- romuscular block: chemical encapsulation of rocuronium bromide by a cyclodex-
dextrins in the pharmaceutical field. Drug Dev Ind Pharm 17:1503–1549. trin-based synthetic host. Angew Chem Int Ed Engl 41:266–270.
Benet LZ and Cummins CL (2001) The drug efflux-metabolism alliance: biochemical Bonini M, Rossi S, Karlsson G, Almgren M, Lo Nostro P, and Baglioni P (2006) Self-
aspects. Adv Drug Deliv Rev 50 (Suppl 1):S3–S11. assembly of beta-cyclodextrin in water. Part 1. Cryo-TEM and dynamic and static
Benet LZ, Cummins CL, and Wu CY (2004) Unmasking the dynamic interplay be- light scattering. Langmuir 22:1478–1484.
tween efflux transporters and metabolic enzymes. Int J Pharm 277:3–9. Borgstrom B, Dahlqvist A, Lundh G, and Sjovall J (1957) Studies of intestinal di-
Berge SM, Bighley LD, and Monkhouse DC (1977) Pharmaceutical salts. J Pharm Sci gestion and absorption in the human. J Clin Invest 36:1521–1536.
66:1–19. Borst P and Elferink RO (2002) Mammalian ABC transporters in health and disease.
Bergeron RJ, Channing MA, and McGovern KA (1978) Dependence of cycloamylose- Annu Rev Biochem 71:537–592.
substrate binding on charge. J Am Chem Soc 100:2878–2883. Bouqet MP and Wagner DR (2012) Injectable formulations of poorly water-soluble
Bergström CAS, Luthman K, and Artursson P (2004) Accuracy of calculated pH- drugs, in Formulating Poorly Water-Soluble Drugs (Williams RO, Watts AB,
dependent aqueous drug solubility. Eur J Pharm Sci 22:387–398. and Miller DA eds) pp 209–242, Springer, New York.
Berk PD and Stump DD (1999) Mechanisms of cellular uptake of long chain free fatty Boyd BJ (2008) Past and future evolution in colloidal drug delivery systems. Expert
acids. Mol Cell Biochem 192:17–31. Opin Drug Deliv 5:69–85.
Berthier L and Biroli G (2011) Theoretical perspective on the glass transition and Boyd BJ (2012) LIPIDS: Towards an understanding of nanostructures formed during
amorphous materials. Rev Mod Phys 83:587–645. digestion, and their impact on drug delivery and absorption. Am Pharmaceutical
Beskid G, Unowsky J, Behl CR, Siebelist J, Tossounian JL, McGarry CM, Shah NH, Rev 15:58.
and Cleeland R (1988) Enteral, oral, and rectal absorption of ceftriaxone using Boyd BJ, Dong Y-D, and Rades T (2009) Nonlamellar liquid crystalline nano-
glyceride enhancers. Chemotherapy 34:77–84. structured particles: advances in materials and structure determination. J Lipo-
Bethune SJ, Huang N, Jayasankar A, and Rodriguez-Hornedo N (2009) Un- some Res 19:12–28.
derstanding and predicting the effect of cocrystal components and pH on cocrystal Branham ML, Moyo T, and Govender T (2012) Preparation and solid-state charac-
solubility. Cryst Growth Des 9:3976–3988. terization of ball milled saquinavir mesylate for solubility enhancement. Eur J
Bettinetti G, Gazzaniga A, Mura P, Giordano F, and Setti M (1992) Thermal-behavior Pharm Biopharm 80:194–202.
and dissolution properties of naproxen in combinations with chemically modified Bras AR, Dionisio M, and Correia NT (2008) Molecular motions in amorphous
beta-cyclodextrins. Drug Dev Ind Pharm 18:39–53. pharmaceuticals, in Complex Systems (Tokuyama M, Oppenheim I, and Nishiyama
478 Williams et al.

H eds) pp 91–96; AIP Conference Proceedings; 2007 September 25–28; Sendai Caliph SM, Faassen WA, Vogel GM, and Porter CJH (2009) Oral bioavailability
Japan. assessment and intestinal lymphatic transport of Org 45697 and Org 46035, two
Bras AR, Merino EG, Neves PD, Fonseca IM, Dionisio M, Schonhals A, and Correia highly lipophilic novel immunomodulator analogues. Curr Drug Deliv 6:359–366.
NT (2011) Amorphous ibuprofen confined in nanostructured silica materials: Caliph SM, Trevaskis NL, Charman WN, and Porter CJH (2012) Intravenous dosing
a dynamical approach. J Phys Chem C 115:4616–4623. conditions may affect systemic clearance for highly lipophilic drugs: implications
Bravo González RC, Huwyler J, Boess F, Walter I, and Bittner B (2004) In vitro for lymphatic transport and absolute bioavailability studies. J Pharmaceutical Sci
investigation on the impact of the surface-active excipients remophor EL, Tween 80 101:3540–3546.
and Solutol HS 15 on the metabolism of midazolam. Biopharm Drug Dispos 25: Cannon BJ, Shi Y, and Gupta P (2008) Emulsions, microemulsions, and lipid-based
37–49. drug delivery systems for drug solubilization and delivery. Part I: parenteral
Brazeau G, Sauberan SL, Gatlin L, Wisniecki P, and Shah J (2011) Effect of particle applications, in Water-Insoluble Drug Formulation (Rong L ed) pp 195–226, Taylor
size of parenteral suspensions on in vitro muscle damage. Pharm Dev Technol 16: and Francis, New York.
591–598. Cannon JB (2008) Chemical and physical stability considerations for lipid-based drug
Breitenbach J (2002) Melt extrusion: from process to drug delivery technology. Eur J formulations. Am Pharmaceutical Rev January/February:132–138.
Pharm Biopharm 54:107–117. Capaccioli S, Paluch M, Prevosto D, Wang LM, and Ngai KL (2012) The many-body
Breitenbach J, Schrof W, and Neumann J (1999) Confocal Raman-spectroscopy: an- nature of relaxation processes in glass-forming systems. J Phys Chem Lett 3:
alytical approach to solid dispersions and mapping of drugs. Pharm Res 16: 735–743.
1109–1113. Capaccioli S, Thayyil MS, and Ngai KL (2008) Critical issues of current research on
Brewster ME, Anderson WR, Estes KS, and Bodor N (1991) Development of aqueous the dynamics leading to glass transition. J Phys Chem B 112:16035–16049.
parenteral formulations for carbamazepine through the use of modified cyclo- Cappello B, di Maio C, Iervolino M, and Miro A (2007) Combined effect of hydroxypropyl
dextrins. J Pharm Sci 80:380–383. methylcellulose and hydroxypropyl-beta-cyclodextrin on physicochemical and dis-
Brewster ME and Loftsson T (2007) Cyclodextrins as pharmaceutical solubilizers. solution properties of celecoxib. J Incl Phenom Macrocycl Chem 59:237–244.
Adv Drug Deliv Rev 59:645–666. Carale TR, Pham QT, and Blankschtein D (1994) Salt effects on intramicellar
Brewster ME, Simpkins JW, Hora MS, Stern WC, and Bodor N (1989) The potential interactions and micellization of nonionic surfactants in aqueous-solutions. Lang-
use of cyclodextrins in parenteral formulations. J Parenter Sci Technol 43:231–240. muir 10:109–121.
Brewster ME, Vandecruys R, Peeters J, Neeskens P, Verreck G, and Loftsson T Carlert S, Pålsson A, Hanisch G, von Corswant C, Nilsson C, Lindfors L, Lennernäs
(2008a) Comparative interaction of 2-hydroxypropyl-beta-cyclodextrin and H, and Abrahamsson B (2010) Predicting intestinal precipitation: a case example
sulfobutylether-beta-cyclodextrin with itraconazole: phase-solubility behavior and for a basic BCS class II drug. Pharm Res 27:2119–2130.
stabilization of supersaturated drug solutions. Eur J Pharm Sci 34:94–103. Carney LG and Hill RM (1976) Human tear pH: diurnal variations. Arch Ophthalmol
Brewster ME, Vandecruys R, Verreck G, Noppe M, and Peeters J (2002) A novel 94:821–824.
cyclodextrin-containing glass thermoplastic system (GTS) for formulating poorly Caron V, Bhugra C, and Pikal MJ (2010) Prediction of onset of crystallization in
water soluble drug candidates: preclinical and clinical results. J Incl Phenom amorphous pharmaceutical systems: phenobarbital, nifedipine/PVP, and pheno-
Macrocycl Chem 44:35–38. barbital/PVP. J Pharm Sci 99:3887–3900.
Brewster ME, Vandecruys R, Verreck G, and Peeters J (2008b) Supersaturating drug Carrier RL, Miller LA, and Ahmed I (2007) The utility of cyclodextrins for enhancing
delivery systems: effect of hydrophilic cyclodextrins and other excipients on the oral bioavailability. J Control Release 123:78–99.
formation and stabilization of supersaturated drug solutions. Pharmazie 63:217–220. Carstensen JT (1995) Drug Stability: Principles and Practices, Marcel Dekker, Inc.,
Bristol-Myers Squibb Company (2007) Prescribing Information Taxol (pacliataxel) New York.
injection. Princeton, NJ. Casley-Smith JR (1962) The identification of chylomicra and lipoproteins in tissue
Brittain HG (2010) Polymorphism and solvatomorphism 2008. J Pharm Sci 99: sections and their passage into jejunal lacteals. J Cell Biol 15:259–277.
3648–3664. Cavallari C, Rodriguez L, Albertini B, Passerini N, Rosetti F, and Fini A (2005)
Brocks DR and Wasan KM (2002) The influence of lipids on stereoselective phar- Thermal and fractal analysis of diclofenac/Gelucire 50/13 microparticles obtained
macokinetics of halofantrine: important implications in food-effect studies in- by ultrasound-assisted atomization. J Pharm Sci 94:1124–1134.
volving drugs that bind to lipoproteins. J Pharm Sci 91:1817–1826. Cavalli R, Bocca C, Miglietta A, Caputo O, and Gasco MR (1999) Albumin adsorption
Broman E, Khoo C, and Taylor LS (2001) A comparison of alternative polymer on stealth and non-stealth solid lipid nanoparticles. STP Pharma Sci 9:183–189.
excipients and processing methods for making solid dispersions of a poorly water Chan HK and Kwok PCL (2011) Production methods for nanodrug particles using the
soluble drug. Int J Pharm 222:139–151. bottom-up approach. Adv Drug Deliv Rev 63:406–416.
Brouwers J, Brewster ME, and Augustijns P (2009) Supersaturating drug delivery Chan KLA and Kazarian SG (2004) FTIR spectroscopic imaging of dissolution of
systems: the answer to solubility-limited oral bioavailability? J Pharm Sci 98: a solid dispersion of nifedipine in poly(ethylene glycol). Mol Pharm 1:331–335.
2549–2572. Chan L, Humma LM, Schriever CA, Fahsingbauer LA, Dominguez CP, and Baum CL
Brouwers J, Tack J, Lammert F, and Augustijns P (2006) Intraluminal drug and (2004) Vitamin E formulation affects digoxin absorption by inhibiting P-
formulation behavior and integration in in vitro permeability estimation: a case glycoprotein (P-gp) in humans. Clin Pharmacol Ther 75:95.
study with amprenavir. J Pharm Sci 95:372–383. Chan OH, Schmid HL, Stilgenbauer LA, Howson W, Horwell DC, and Stewart BH
Bruce CD, Fegely KA, Rajabi-Siahboomi AR, and McGinity JW (2010) The influence (1998) Evaluation of a targeted prodrug strategy of enhance oral absorption of
of heterogeneous nucleation on the surface crystallization of guaifenesin from melt poorly water-soluble compounds. Pharm Res 15:1012–1018.
extrudates containing Eudragit L10055 or Acryl-EZE. Eur J Pharm Biopharm 75: Chandler D (2005) Interfaces and the driving force of hydrophobic assembly. Nature
71–78. 437:640–647.
Brunham LR, Kruit JK, Iqbal J, Brunham LR, Kruit JK, Iqbal J, Fievet C, Timmins JM, Chang RK and Shojaei AH (2004) Effect of a lipoidic excipient on the absorption
Pape TD, Coburn BA, Bissada N, Staels B, and Groen AK, et al. (2006) Intestinal profile of compound UK 81252 in dogs after oral administration. J Pharm Pharm
ABCA1 directly contributes to HDL biogenesis in vivo. J Clin Invest 116: Sci 7:8–12.
1052–1062. Chang SW, Reddy V, Pereira T, Dean BJ, Xia YQ, Seto C, Franklin RB, and Karanam BV
Brunner LJ and Bai S (2000) Effect of dietary oil intake on hepatic cytochrome P450 (2007) The pharmacokinetics and disposition of MK-0524, a prosglandin D2 re-
activity in the rat. J Pharm Sci 89:1022–1027. ceptor 1 antagonist, in rats, dogs and monkeys. Xenobiotica 37:514–533.
Buggins TR, Dickinson PA, and Taylor G (2007) The effects of pharmaceutical excip- Chang T, Benet LZ, and Hebert MF (1996) The effect of water-soluble vitamin E on
ients on drug disposition. Adv Drug Deliv Rev 59:1482–1503. cyclosporine pharmacokinetics in healthy volunteers. Clin Pharmacol Ther 59:
Bummer PM (2004) Physical chemical considerations of lipid-based oral drug de- 297–303.
livery: solid lipid nanoparticles. Crit Rev Ther Drug Carrier Syst 21:1–20. Charman SA, Charman WN, Rogge MC, Wilson TD, Dutko FJ, and Pouton CW
Bunjes H (2010) Lipid nanoparticles for the delivery of poorly water-soluble drugs. (1992) Self-emulsifying drug delivery systems: formulation and biopharmaceutic
J Pharm Pharmacol 62:1637–1645. evaluation of an investigational lipophilic compound. Pharm Res 9:87–93.
Burcham DL, Aungst BA, Hussain M, Gorko MA, Quon CY, and Huang SM (1995) Charman SA, Perry CS, Chiu FCK, McIntosh KA, Prankerd RJ, and Charman WN
The effect of absorption enhancers on the oral absorption of the GP IIB/IIIA re- (2006) Alteration of the intravenous pharmacokinetics of a synthetic ozonide an-
ceptor antagonist, DMP 728, in rats and dogs. Pharm Res 12:2065–2070. timalarial in the presence of a modified cyclodextrin. J Pharm Sci 95:256–267.
Buschmann HJ, Jansen K, and Schollmeyer E (2000) Cucurbituril and alpha- and Charman WN, Rogge MC, Boddy AW, and Berger BM (1993) Effect of food and
beta-cyclodextrins as ligands for the complexation of nonionic surfactants and a monoglyceride emulsion formulation on danazol bioavailability. J Clin Phar-
polyethyleneglycols in aqueous solutions. J Incl Phenom Macrocycl Chem 37: macol 33:381–386.
231–236. Charman WN and Stella VJ (1986) Effect of lipid class and lipid vehicle volume on
Butler JM and Dressman JB (2010) The developability classification system: appli- the intestinal lymphatic transport of DDT. Int J Pharm 33:165–172.
cation of biopharmaceutics concepts to formulation development. J Pharm Sci 99: Chaubal MV (2004) Application of drug delivery technologies in lead candidate se-
4940–4954. lection and optimization. Drug Discov Today 9:603–609.
Byrn SR, Stowell JG, and Pfeiffer R (1999) Solid State Chemistry of Drugs, SSCI Chauhan B, Shimpi S, and Paradkar A (2005) Preparation and characterization of
Press, West Lafayette, IN. etoricoxib solid dispersions using lipid carriers by spray drying technique. AAPS
Byrn SR, Zografi G, and Chen XS (2010) Accelerating proof of concept for small PharmSciTech 6:E405–E412.
molecule drugs using solid-state chemistry. J Pharm Sci 99:3665–3675. Chemburkar SR, Bauer J, Deming K, Spiwek H, Patel K, Morris J, Henry R, Spanton
Cabral Marques HM, Hadgraft J, and Kellaway IW (1990) Studies of cyclodextrin S, Dziki W, and Porter W, et al. (2000) Dealing with the impact of ritonavir poly-
inclusion complexes. 1. The salbutamol-cyclodextrin complex as studied by phase morphs on the late stages of bulk drug process development. Org Process Res Dev 4:
solubility and Dsc. Int J Pharm 63:259–266. 413–417.
Cal K and Sollohub K (2010) Spray drying technique. I: Hardware and process Chen J, Ping QE, Guo JX, Liu ML, and Cai BC (2007) Effect of injection routes on
parameters. J Pharm Sci 586. pharmacokinetics and lactone/carboxylate equilibrium of 9-nitrocamptothecin in
Caliph SM, Charman WN, and Porter CJH (2000) Effect of short-, medium-, and long- rats. Int J Pharm 340:29–33.
chain fatty acid-based vehicles on the absolute oral bioavailability and intestinal Chen LJ, Lin SY, and Huang CC (1998) Effect of hydrophobic chain length of sur-
lymphatic transport of halofantrine and assessment of mass balance in lymph- factants on enthalpy-entropy compensation of micellization. J Phys Chem B 102:
cannulated and non-cannulated rats. J Pharm Sci 89:1073–1084. 4350–4356.
Strategies for Low Drug Solubility 479
Chen ML (2008) Lipid excipients and delivery systems for pharmaceutical de- Clement M, Pugh W, and Parikh I (1992) Tissue distribution and plasma clearance of
velopment: a regulatory perspective. Adv Drug Deliv Rev 60:768–777. a novel microcrystalline-coated flurbiprofen formulation. Pharmacologist 34:204.
Chen ML, Amidon GL, Benet LZ, Lennernas H, and Yu LX (2011) The BCS, BDDCS, Cohen DE and Carey MC (1990) Rapid (1 hour) high performance gel filtration
and regulatory guidances. Pharm Res 28:1774–1778. chromatography resolves coexisting simple micelles, mixed micelles, and vesicles
Chen X, Benhayoune Z, Williams RO, and Johnston KP (2004a) Rapid dissolution of in bile. J Lipid Res 31:2103–2112.
high potency itraconazole particles produced by evaporative precipitation into Cole ET, Cadé D, and Benameur H (2008) Challenges and opportunities in the en-
aqueous solution. J Drug Delivery Sci Technol 14:299–304. capsulation of liquid and semi-solid formulations into capsules for oral adminis-
Chen XQ, Antman MD, Gesenberg C, and Gudmundsson OS (2006) Discovery tration. Adv Drug Deliv Rev 60:747–756.
pharmaceutics: challenges and opportunities. AAPS J 8:E402–E408. Colombo T, Gonzalez Paz O, Zucchetti M, Maneo A, Sessa C, Goldhirsch A,
Chen XX, Young TJ, Sarkari M, Williams RO 3rd, and Johnston KP (2002) Prepa- and D’Incalci M (1996) Paclitaxel induces significant changes in epidoxorubicin
ration of cyclosporine A nanoparticles by evaporative precipitation into aqueous distribution in mice. Ann Oncol 7:801–805.
solution. Int J Pharm 242:3–14. Colombo T, Parisi I, Zucchetti M, Sessa C, Goldhirsch A, and D’Incalci M (1999)
Chen Y, Zhang GGZ, Neilly J, Marsh K, Mawhinney D, and Sanzgiri YD (2004b) Pharmacokinetic interactions of paclitaxel, docetaxel and their vehicles with
Enhancing the bioavailability of ABT-963 using solid dispersion containing Pluronic doxorubicin. Ann Oncol 10:391–395.
F-68. Int J Pharm 286:69–80. Constantinides PP, Chaubal MV, and Shorr R (2008) Advances in lipid nanodisper-
Chen YM, Lin PC, Tang M, and Chen YP (2010) Solid solubility of antilipemic agents sions for parenteral drug delivery and targeting. Adv Drug Deliv Rev 60:757–767.
and micronization of gemfibrozil in supercritical carbon dioxide. J Supercrit Fluids Constantinides PP, Scalart JP, Lancaster C, Marcello J, Marks G, Ellens H,
52:175–182. and Smith PL (1994) Formulation and intestinal absorption enhancement evalu-
Cheney ML, Shan N, Healey ER, Hanna M, Wojtas L, Zaworotko MJ, Sava V, Song SJ, ation of water-in-oil microemulsions incorporating medium-chain glycerides.
and Sanchez-Ramos JR (2010) Effects of crystal form on solubility and pharma- Pharm Res 11:1385–1390.
cokinetics: a crystal engineering case study of lamotrigine. Cryst Growth Des 10: Constantinides PP and Wasan KM (2007) Lipid formulation strategies for enhancing
394–405. intestinal transport and absorption of P-glycoprotein (P-gp) substrate drugs: in
Cheney ML, Weyna DR, Shan N, Hanna M, Wojtas L, and Zaworotko MJ (2011) vitro/in vivo case studies. J Pharm Sci 96:235–248.
Coformer selection in pharmaceutical cocrystal development: a case study of Constantinides PP, Welzel G, Ellens H, Smith PL, Sturgis S, Yiv SH, and Owen AB
a meloxicam aspirin cocrystal that exhibits enhanced solubility and pharmacoki- (1996) Water-in-oil microemulsions containing medium-chain fatty acids/salts:
netics. J Pharm Sci 100:2172–2181. formulation and intestinal absorption enhancement evaluation. Pharm Res 13:
Cherukuvada S, Babu NJ, and Nangia A (2011) Nitrofurantoin-p-aminobenzoic acid 210–215.
cocrystal: hydration stability and dissolution rate studies. J Pharm Sci 100: Coon JS, Knudson W, Clodfelter K, Lu B, and Weinstein RS (1991) Solutol HS 15,
3233–3244. nontoxic polyoxyethylene esters of 12-hydroxystearic acid, reverses multidrug re-
Chiang PC, South SA, Daniels JS, Anderson DR, Wene SP, Albin LA, Mourey RJ, sistance. Cancer Res 51:897–902.
and Selbo JG (2009) Aqueous versus non-aqueous salt delivery strategies to en- Cornaire G, Woodley J, Hermann P, Cloarec A, Arellano C, and Houin G (2004)
hance oral bioavailability of a mitogen-activated protein kinase-activated protein Impact of excipients on the absorption of P-glycoprotein substrates in vitro and in
kinase (MK-2) inhibitor in rats. J Pharm Sci 98:248–256. vivo. Int J Pharm 278:119–131.
Chiappetta DA and Sosnik A (2007) Poly(ethylene oxide)-poly(propylene oxide) block Corrigan OI (1986) Retardation of polymeric carrier dissolution by dispersed drugs:
copolymer micelles as drug delivery agents: improved hydrosolubility, stability and factors influencing the dissolution of solid dispersions containing polyethylene
bioavailability of drugs. Eur J Pharm Biopharm 66:303–317. glycols. Drug Dev Ind Pharm 12:1777–1793.
Chiarella RA, Davey RJ, and Peterson ML (2007) Making co-crystals: the utility of Corrigan OI (2007) Salt forms: pharmaceutical aspects, in Encyclopedia of Pharma-
ternary phase diagrams. Cryst Growth Des 7:1223–1226. ceutical Technology (Swarbrick J ed) pp 3177–3188, Informa Healthcare, New
Chien YW (1981) Long-acting parenteral drug formulations. J Parenter Sci Technol York.
35:106–139. Corrigan OI and Crean AM (2002) Comparative physicochemical properties of
Childs SL, Chyall LJ, Dunlap JT, Smolenskaya VN, Stahly BC, and Stahly GP (2004) hydrocortisone-PVP composites prepared using supercritical carbon dioxide by the
Crystal engineering approach to forming cocrystals of amine hydrochlorides with GAS anti-solvent recrystallization process, by coprecipitation and by spray drying.
organic acids: molecular complexes of fluoxetine hydrochloride with benzoic, suc- Int J Pharm 245:75–82.
cinic, and fumaric acids. J Am Chem Soc 126:13335–13342. Corrigan OI and Stanley CT (1982) Mechanism of drug dissolution rate enhancement
Childs SL and Hardcastle KI (2007) Cocrystals of piroxicam with carboxylic acids. from beta-cyclodextrin-drug systems. J Pharm Pharmacol 34:621–626.
Cryst Growth Des 7:1291–1304. Cotton ML, Lamarche P, Motola S, and Vadas EB (1994) L-649,923—the selection of
Childs SL, Rodriguez-Hornedo N, Reddy LS, Jayasankar A, Maheshwari C, an appropriate salt form and preparation of a stable oral formulation. Int J Pharm
McCausland L, Shipplett R, and Stahly BC (2008) Screening strategies based on 109:237–249.
solubility and solution composition generate pharmaceutically acceptable cocrys- Cox JW, Sage GP, Wynalda MA, Ulrich RG, Larson PG, and Su CC (1991) Plasma
tals of carbamazepine. Cryst Eng Comm 10:856–864. compatibility of injectables: comparison of intravenous U-74006F, a 21-amino-
Childs SL and Zaworotko MJ (2009) The reemergence of cocrystals: the crystal clear steroid antioxidant, with dilantin brand of parenteral phenytoin. J Pharm Sci 80:
writing is on the wall introduction to virtual special issue on pharmaceutical 371–375.
cocrystals. Cryst Growth Des 9:4208–4211. Craig DQM (1990) Polyethyelene glycols and drug release. Drug Dev Ind Pharm 16:
Chingunpitak J, Puttipipatkhachorn S, Tozuka Y, Moribe K, and Yamamoto K (2008) 2501–2526.
Micronization of dihydroartemisinin by rapid expansion of supercritical solutions. Craig DQM (1992) Applications of low-frequency dielectric-spectroscopy to the
Drug Dev Ind Pharm 34:609–617. pharmaceutical sciences. Drug Dev Ind Pharm 18:1207–1223.
Chiou WL and Riegelman S (1970) Oral absorption of griseofulvin in dogs: increased Craig DQM (2002) The mechanisms of drug release from solid dispersions in water-
absorption via solid dispersion in polyethylene glycol 6000. J Pharm Sci 59: soluble polymers. Int J Pharm 231:131–144.
937–942. Craig DQM and Newton JM (1992) The dissolution of nortriptyline HCl from
Chiou WL and Riegelman S (1971) Pharmaceutical applications of solid dispersion polyethylene-glycol solid dispersions. Int J Pharm 78:175–182.
systems. J Pharm Sci 60:1281–1302. Crew MD, Friesen DT, Hancock BC, Macri C, Nightingale JA, and Shanker R (2007),
Choo EF, Alicke B, Boggs J, Dinkel V, Gould S, Grina J, West K, Menghrajani K, Ran YQ, inventors; Pfizer, Inc., assignee. Pharmaceutical compositions of a sparingly solu-
and Rudolph J, et al. (2011) Preclinical assessment of novel BRAF inhibitors: in- ble glycogen phosphorylase inhibitor U.S. patent 7,235,260. 2007 June 26.
tegrating pharmacokinetic-pharmacodynamic modelling in the drug discovery Cromwell WC, Bystrom K, and Eftink MR (1985) Cyclodextrin adamantanecarboxylate
process. Xenobiotica 41:1076–1087. inclusion complexes: studies of the variation in cavity size. J Phys Chem 89:
Chow K, Tong HHY, Lum S, and Chow AHL (2008) Engineering of pharmaceutical 326–332.
materials: an industrial perspective. J Pharm Sci 97:2855–2877. Crounse RG (1961) Human pharmacology of griseofulvin: the effect of fat intake on
Chow SL and Hollander D (1979) A dual, concentration-dependent absorption gastrointestinal absorption. J Invest Dermatol 37:529–533.
mechanism of linoleic acid by rat jejunum in vitro. J Lipid Res 20:349–356. Crowley KJ and Zografi G (2001) The use of thermal methods for predicting glass-
Chowdary KPR and Srinivas SV (2006) Influence of hydrophilic polymers on former fragility. Thermochim Acta 380:79–93.
celecoxib complexation with hydroxypropyl beta-cyclodextrin. AAPS PharmSciTech Crowley MM, Zhang F, Repka MA, Thumma S, Upadhye SB, Battu SK, McGinity
7:79. JW, and Martin C (2007) Pharmaceutical applications of hot-melt extrusion. Part I.
Chowhan ZT (1978) pH-solubility profiles or organic carboxylic acids and their salts. Drug Dev Ind Pharm 33:909–926.
J Pharm Sci 67:1257–1260. Croy SR and Kwon GS (2004) The effects of pluronic block copolymers on the ag-
Christensen JO, Schultz K, Mollgaard B, Kristensen HG, and Mullertz A (2004) gregation state of nystatin. J Control Release 95:161–171.
Solubilisation of poorly water-soluble drugs during in vitro lipolysis of medium- Cuiné JF, Charman WN, Pouton CW, Edwards GA, and Porter CJH (2007) In-
and long-chain triacylglycerols. Eur J Pharm Sci 23:287–296. creasing the proportional content of surfactant (Cremophor EL) relative to lipid in
Christensen KL, Pedersen GP, and Kristensen HG (2001) Technical optimisation of self-emulsifying lipid-based formulations of danazol reduces oral bioavailability in
redispersible dry emulsions. Int J Pharm 212:195–202. beagle dogs. Pharm Res 24:748–757.
Christiansen A, Backensfeld T, Denner K, and Weitschies W (2011) Effects of non- Cuiné JF, McEvoy CL, Charman WN, Pouton CW, Edwards GA, Benameur H,
ionic surfactants on cytochrome P450-mediated metabolism in vitro. Eur J Pharm and Porter CJH (2008) Evaluation of the impact of surfactant digestion on the
Biopharm 78:166–172. bioavailability of danazol after oral administration of lipidic self-emulsifying for-
Christiansen A, Backensfeld T, and Weitschies W (2010) Effects of non-ionic sur- mulations to dogs. J Pharm Sci 97:995–1012.
factants on in vitro triglyceride digestion and their susceptibility to digestion by Curatolo W (1998) Physical chemical properties of oral drug candidates in the dis-
pancreatic enzymes. Eur J Pharm Sci 41:376–382. covery and exploratory development settings. Pharm Sci Technol Today 1:387–393.
Chung PH, Chin TF, and Lach JL (1970) Kinetics of the hydrolysis of pilocarpine in Curatolo W, Nightingale JA, and Herbig SM (2009) Utility of hydroxypropylmethylcellulose
aqueous solution. J Pharm Sci 59:1300–1306. acetate succinate (HPMCAS) for initiation and maintenance of drug supersatura-
Cirri M, Maestrelli F, Corti G, Furlanetto S, and Mura P (2006) Simultaneous effect tion in the GI milieu. Pharm Res 26:1419–1431.
of cyclodextrin complexation, pH, and hydrophilic polymers on naproxen solubili- D’Addio SM and Prud’homme RK (2011) Controlling drug nanoparticle formation by
zation. J Pharm Biomed Anal 42:126–131. rapid precipitation. Adv Drug Deliv Rev 63:417–426.
480 Williams et al.

Dahan A, Duvdevani R, Shapiro I, Elmann A, Finkelstein E, and Hoffman A (2008) Di L, Fish PV, and Mano T (2012) Bridging solubility between drug discovery and
The oral absorption of phospholipid prodrugs: in vivo and in vitro mechanistic development. Drug Discov Today 17:486–495.
investigation of trafficking of a lecithin-valproic acid conjugate following oral ad- Di L, Kerns EH, and Carter GT (2009) Drug-like property concepts in pharmaceutical
ministration. J Control Release 126:1–9. design. Curr Pharm Des 15:2184–2194.
Dahan A and Hoffman A (2005) Evaluation of a chylomicron flow blocking approach Dias MMR, Raghavan SL, Pellett MA, and Hadgraft J (2003) The effect of beta-
to investigate the intestinal lymphatic transport of lipophilic drugs. Eur J Pharm cyclodextrins on the permeation of diclofenac from supersaturated solutions. Int J
Sci 24:381–388. Pharm 263:173–181.
Dahan A and Hoffman A (2006) Use of a dynamic in vitro lipolysis model to ratio- di Cagno M, Stein PC, Skalko-Basnet N, Brandl M, and Bauer-Brandl A (2011)
nalize oral formulation development for poor water soluble drugs: correlation with Solubilization of ibuprofen with b-cyclodextrin derivatives: energetic and struc-
in vivo data and the relationship to intra-enterocyte processes in rats. Pharm Res tural studies. J Pharm Biomed Anal 55:446–451.
23:2165–2174. Di Maio S and Carrier RL (2011) Gastrointestinal contents in fasted state and post-
Dahan A and Hoffman A (2007) The effect of different lipid based formulations on the lipid ingestion: in vivo measurements and in vitro models for studying oral drug
oral absorption of lipophilic drugs: the ability of in vitro lipolysis and consecutive ex delivery. J Control Release 151:110–122.
vivo intestinal permeability data to predict in vivo bioavailability in rats. Eur J Dintaman JM and Silverman JA (1999) Inhibition of P-glycoprotein by D-alpha-
Pharm Biopharm 67:96–105. tocopheryl polyethylene glycol 1000 succinate (TPGS). Pharm Res 16:1550–1556.
Dahan A, Miller JM, and Amidon GL (2009) Prediction of solubility and permeability DiNunzio JC, Brough C, Miller DA, Williams RO 3rd, and McGinity JW (2010a)
class membership: provisional BCS classification of the world’s top oral drugs. Applications of KinetiSol dispersing for the production of plasticizer free amor-
AAPS J 11:740–746. phous solid dispersions. Eur J Pharm Sci 40:179–187.
Dahan A, Miller JM, Hoffman A, Amidon GE, and Amidon GL (2010) The solubility- DiNunzio JC, Brough C, Miller DA, Williams RO 3rd, and McGinity JW (2010b)
permeability interplay in using cyclodextrins as pharmaceutical solubilizers: Fusion processing of itraconazole solid dispersions by kinetisol dispersing: a com-
mechanistic modeling and application to progesterone. J Pharm Sci 99:2739–2749. parative study to hot melt extrusion. J Pharm Sci 99:1239–1253.
Dai WG (2010) In vitro methods to assess drug precipitation. Int J Pharm 393:1–16. Dixit RP and Nagarsenker MS (2010) Optimized microemulsions and solid micro-
Dai WG, Dong LC, Li S, and Deng ZY (2008a) Combination of pluronic/vitamin emulsion systems of simvastatin: characterization and in vivo evaluation. J Pharm
E TPGS as a potential inhibitor of drug precipitation. Int J Pharm 355:31–37. Sci 99:4892–4902.
Dai WG, Dong LC, Shi XF, Nguyen J, Evans J, Xu YD, and Creasey AA (2007) Dixon BJ (2010) Mechanisms of chylomicron uptake into lacteals. Ann NY Acad Sci 1:
Evaluation of drug precipitation of solubility-enhancing liquid formulations using E52–E57.
milligram quantities of a new molecular entity (NME). J Pharm Sci 96:2957–2969. Dixon JB, Raghunathan S, and Swartz MA (2009) A tissue-engineered model of the
Dai WG, Pollock-Dove C, Dong LC, and Li S (2008b) Advanced screening assays to intestinal lacteal for evaluating lipid transport by lymphatics. Biotechnol Bioeng
rapidly identify solubility-enhancing formulations: high-throughput, miniaturiza- 103:1224–1235.
tion and automation. Adv Drug Deliv Rev 60:657–672. Do TT, Van Speybroeck M, Mols R, Annaert P, Martens J, Van Humbeeck J, Vermant
Damian F, Blaton N, Kinget R, and Van den Mooter G (2002) Physical stability of J, Augustijns P, and Van den Mooter G (2011) The conflict between in vitro release
solid dispersions of the antiviral agent UC-781 with PEG 6000, Gelucire 44/14 and studies in human biorelevant media and the in vivo exposure in rats of the lipo-
PVP K30. Int J Pharm 244:87–98. philic compound fenofibrate. Int J Pharm 414:118–124.
Dannenfelser RM, Surakitbanharn Y, Tabibi SE, and Yalkowsky SH (1996) Parenteral Dobbins WO 3rd (1971) Intestinal mucosal lacteal in transport of macromolecules
formulation of flavopiridol (NSC-649890). PDA J Pharm Sci Technol 50:356–359. and chylomicrons. Am J Clin Nutr 24:77–90.
Danson S, Ferry D, Alakhov V, Margison J, Kerr D, Jowle D, Brampton M, Halbert G, Dobbins WO 3rd and Rollins EL (1970) Intestinal mucosal lymphatic permeability:
and Ranson M (2004) Phase I dose escalation and pharmacokinetic study of an electron microscopic study of endothelial vesicles and cell junctions. J Ultra-
pluronic polymer-bound doxorubicin (SP1049C) in patients with advanced cancer. struct Res 33:29–59.
Br J Cancer 90:2085–2091. Dobbs JM, Wong JM, Lahiere RJ, and Johnston KP (1987) Modification of su-
Dantuluri AKR, Amin A, Puri V, and Bansal AK (2011) Role of a-relaxation on percritical fluid phase-behavior using polar cosolvents. Ind Eng Chem Res 26:
crystallization of amorphous celecoxib above Tg probed by dielectric spectroscopy. 56–65.
Mol Pharm 8:814–822. Doherty C and York P (1987) Mechanisms of dissolution of frusemide/PVP solid
Date AA and Nagarsenker MS (2008) Parenteral microemulsions: an overview. Int J dispersions. Int J Pharm 34:197–205.
Pharm 355:19–30. Dollo G, Le Corre P, Chollet M, Chevanne F, Bertault M, Burgot JL, and Le Verge R
Datta S and Grant DJW (2004) Crystal structures of drugs: advances in de- (1999) Improvement in solubility and dissolution rate of 1, 2-dithiole-3-thiones
termination, prediction and engineering. Nat Rev Drug Discov 3:42–57. upon complexation with beta-cyclodextrin and its hydroxypropyl and sulfobutyl
David SE, Timmins P, and Conway BR (2012) Impact of the counterion on the sol- ether-7 derivatives. J Pharm Sci 88:889–895.
ubility and physicochemical properties of salts of carboxylic acid drugs. Drug Dev Dollo G, Le Corre P, Guérin A, Chevanne F, Burgot JL, and Leverge R (2003) Spray-
Ind Pharm 38:93–103. dried redispersible oil-in-water emulsion to improve oral bioavailability of poorly
Davies RO and Jones GO (1953) Thermodynamic and kinetic properties of glasses. soluble drugs. Eur J Pharm Sci 19:273–280.
Adv Phys 2:370–410. Dong YC, Ng WK, Surana U, and Tan RBH (2008) Solubilization and preformulation
Davis JM, Matalon L, Watanabe MD, and Blake L (1994) Depot antipsychotic drugs: of poorly water soluble and hydrolysis susceptible N-epoxymethyl-1,8-
place in therapy. Drugs 47:741–773. naphthalimide (ENA) compound. Int J Pharm 356:130–136.
Davis ME and Brewster ME (2004) Cyclodextrin-based pharmaceutics: past, present Dontireddy R and Crean AM (2011) A comparative study of spray-dried and freeze-
and future. Nat Rev Drug Discov 3:1023–1035. dried hydrocortisone/polyvinyl pyrrolidone solid dispersions. Drug Dev Ind Pharm
Davis WW, Pfeiffer RR, and Quay JF (1970) Normal and promoted gastrointestinal 37:1141–1149.
absorption of water-soluble substances. I. Induced rapidly reversible hyper- Dordunoo SK, Ford JL, and Rubinstein MH (1991) Preformulation studies on solid
absorptive state in the canine fundic stomach pouch. J Pharm Sci 59:960–963. dispersions containing triamterene or temazepam in polyethylene glycols or
Day JPR, Rago G, Domke KF, Velikov KP, and Bonn M (2010) Label-free imaging of Gelucire 44/14 for liquid filling of hard gelatin capsules. Drug Dev Ind Pharm 17:
lipophilic bioactive molecules during lipid digestion by multiplex coherent anti- 1685–1713.
Stokes Raman scattering microspectroscopy. J Am Chem Soc 132:8433–8439. Doshi DH, Ravis WR, and Betageri GV (1997) Carbamazepine and polyethylene
De Bie AT, Van Ommen B, and Bär A (1998) Disposition of [14C]gamma-cyclodextrin glycol solid dispersions: preparation, in vitro dissolution, and characterization.
in germ-free and conventional rats. Regul Toxicol Pharmacol 27:150–158. Drug Dev Ind Pharm 23:1167–1176.
de Garavilla L, Liversidge EM, and Liversidge GG (1998), inventors; Nanosystems Douglas SJ, Davis SS, and Illum L (1987) Nanoparticles in drug delivery. Crit Rev
LLC, assignee. Reduction of intravenously administered nanoparticulate- Ther Drug Carrier Syst 3:233–261.
formulation-induced adverse physiological reactions. US Patent 5,835,025. 1998 Dressman JB, Amidon GL, Reppas C, and Shah VP (1998) Dissolution testing as
Nov 10. a prognostic tool for oral drug absorption: immediate release dosage forms. Pharm
de Garavilla L, Peltier N, and Merisko-Liversidge E (1996) Controlling the acute Res 15:11–22.
hemodynamic effects associated with IV administration of particulate drug dis- Dressman JB, Vertzoni M, Goumas K, and Reppas C (2007) Estimating drug solu-
persions in dogs. Drug Dev Res 37:86–96. bility in the gastrointestinal tract. Adv Drug Deliv Rev 59:591–602.
D’Errico G, Ciccarelli D, and Ortona O (2005) Effect of glycerol on micelle formation Driscoll DF (2006a) Lipid injectable emulsions: 2006. Nutr Clin Pract 21:381–386.
by ionic and nonionic surfactants at 25 degrees C. J Colloid Interface Sci 286: Driscoll DF (2006b) Lipid injectable emulsions: pharmacopeial and safety issues.
747–754. Pharm Res 23:1959–1969.
Desai PR, Jain NJ, Sharma RK, and Bahadur P (2001) Effect of additives on the Dubois JL and Ford JL (1985) Similarities in the release rates of different drugs from
micellization of PEO/PPO/PEO block copolymer F127 in aqueous solution. Colloids polyethylene glycol 6000 solid dispersions. J Pharm Pharmacol 37:494–495.
and Surfaces Physicochem Eng Aspects 178:57–69. Edwards GA, Porter CJH, Caliph SM, Khoo SM, and Charman WN (2001) Animal
Descamps M, Willart JF, Dudognon E, and Caron V (2007) Transformation of models for the study of intestinal lymphatic drug transport. Adv Drug Deliv Rev 50:
pharmaceutical compounds upon milling and comilling: the role of T(g). J Pharm 45–60.
Sci 96:1398–1407. Egan TD, Kern SE, Johnson KB, and Pace NL (2003) The pharmacokinetics and
Desiraju GR (2001) Chemistry beyond the molecule. Nature 412:397–400. pharmacodynamics of propofol in a modified cyclodextrin formulation (Captisol)
Desiraju GR (2010) Crystal engineering: a brief overview. J Chem Sci 122:667–675. versus propofol in a lipid formulation (Diprivan): an electroencephalographic and
Desjardins P, Black P, Papageorge M, Norwood T, Shen DD, Norris L, and Ardia A hemodynamic study in a porcine model. Anesth Analg 97:72–79.
(2002) Ibuprofen arginate provides effective relief from postoperative dental pain Ehninger G, Schuler U, Renner U, Ehrsam M, Zeller KP, Blanz J, Storb R, and Deeg HJ
with a more rapid onset of action than ibuprofen. Eur J Clin Pharmacol 58: (1995) Use of a water-soluble busulfan formulation: pharmacokinetic studies in
387–394. a canine model. Blood 85:3247–3249.
de Waard H, Frijlink HW, and Hinrichs WLJ (2011) Bottom-up preparation tech- Eicke HF (1980) Surfactants in nonpolar solvents: aggregation and micellization. Top
niques for nanocrystals of lipophilic drugs. Pharm Res 28:1220–1223. Curr Chem 87:85–145.
de Waard H, Hinrichs WLJ, and Frijlink HW (2008) A novel bottom-up process to Elamin AA, Ahlneck C, Alderborn G, and Nystrom C (1994) Increased metastable
produce drug nanocrystals: controlled crystallization during freeze-drying. J Con- solubility of milled griseofulvin, depending on the formation of a disordered
trol Release 128:179–183. surface-structure. Int J Pharm 111:159–170.
Strategies for Low Drug Solubility 481
Eldem T, Speiser P, and Hincal A (1991) Optimization of spray-dried and -congealed Fernandes CM, Teresa Vieira M, and Veiga FJB (2002) Physicochemical character-
lipid micropellets and characterization of their surface morphology by scanning ization and in vitro dissolution behavior of nicardipine-cyclodextrins inclusion
electron microscopy. Pharm Res 8:47–54. compounds. Eur J Pharm Sci 15:79–88.
Elder DP, Delaney E, Teasdale A, Eyley S, Reif VD, Jacq K, Facchine KL, Oestrich RS, Fernandez S, Chevrier S, Ritter N, Mahler B, Demarne F, Carrière F, and Jannin V
Sandra P, and David F (2010) The utility of sulfonate salts in drug development. J (2009) In vitro gastrointestinal lipolysis of four formulations of piroxicam and
Pharm Sci 99:2948–2961. cinnarizine with the self emulsifying excipients Labrasol and Gelucire 44/14.
El-Eini DID, Barry BW, and Rhodes CT (1976) Micellar size, shape, and hydration of Pharm Res 26:1901–1910.
long-chain polyoxyethylene nonionic surfactants. J Colloid Interface Sci 54: Fernandez S, Jannin V, Rodier JD, Ritter N, Mahler B, and Carriere F (2007)
348–351. Comparative study on digestive lipase activities on the self emulsifying excipient
Elhasi S, Astaneh R, and Lavasanifar A (2007) Solubilization of an amphiphilic drug by Labrasol (R), medium chain glycerides and PEG esters. Biochimica et Biophysica
poly(ethylene oxide)-block-poly(ester) micelles. Eur J Pharm Biopharm 65:406–413. Acta—Mol Cell Biol Lipids 1771:633–640.
Eli WJ, Chen WH, and Xue QJ (2000) Determination of association constants of Fernandez S, Rodier JD, Ritter N, Mahler B, Demarne F, Carrière F, and Jannin V
cyclodextrin-nonionic surfactant inclusion complexes by a partition coefficient (2008) Lipolysis of the semi-solid self-emulsifying excipient Gelucire 44/14 by di-
method. J Incl Phenom Macrocycl Chem 38:37–43. gestive lipases. Biochim Biophys Acta (Amsterdam) 1781:367–375.
Ellis AG, Crinis NA, and Webster LK (1996) Inhibition of etoposide elimination in the Fini A, Fazio G, and Feroci G (1995) Solubility and solubilization properties of non-
isolated perfused rat liver by Cremophor EL and Tween 80. Cancer Chemother steroidal anti-inflammatory drugs. Int J Pharm 126:95–102.
Pharmacol 38:81–87. Fini A, Fazio G, Gonzalez-Rodriguez M, Cavallari C, Passerini N, and Rodriguez L
El-Sayed AAA and Repta AJ (1983) Solubilization and stabilization of an in- (1999) Formation of ion-pairs in aqueous solutions of diclofenac salts. Int J Pharm
vestigational antineoplastic drug (NSC No 278214) in an intravenous formulation 187:163–173.
using an emulsion vehicle. Int J Pharm 13:303–312. Fini A, Fazio G, Hervas MJF, Holgado MA, and Rabasco AM (1996) Factors gov-
Elworthy PH, Florence AT, and Macfarlane CB (1968) Solubilization by Surface erning the dissolution of diclofenac salts. Eur J Pharm Sci 4:231–238.
Active Agents and Its Applications in Chemistry and Biological Sciences, Chapman Fini A, Rodriguez L, Cavallari C, Albertini B, and Passerini N (2002) Ultrasound-
and Hall, London. compacted and spray-congealed indomethacin/polyethyleneglycol systems. Int J
Elworthy PH and Patel MS (1982) Demonstration of maximum solubilization in Pharm 247:11–22.
a polyoxyethylene alkyl ether series of non-ionic surfactants. J Pharm Pharmacol Fleischman SG, Kuduva SS, McMahon JA, Moulton B, Walsh RDB, Rodriguez-
34:543–546. Hornedo N, and Zaworotko MJ (2003) Crystal engineering of the composition of
Elworthy PH and Treon JF (1966) Physiological activity of nonionic surfactants, in pharmaceutical phases: multiple-component crystalline solids involving carbamaz-
Nonionic Surfactants (Schick JM ed) pp 923–970, Marcel Dekker, New York. epine. Cryst Growth Des 3:909–919.
European Medicines Agency (EMEA); (2007) CHMP Assessment Report for Viracept, Fleisher D, Bong R, and Stewart BH (1996) Improved oral drug delivery: solubility
EMEA/CHMP/492059, 20 September 2007. Committee for Medicinal Products for limitations overcome by the use of prodrugs. Adv Drug Deliv Rev 19:115–130.
Human Use (CHMP), London. Flory PJ (1953) Principles of Polymer Chemistry, Cornell University Press, Ithaca, NY.
Endemann G, Stanton LW, Madden KS, Bryant CM, White RT, and Protter AA Flourié B, Molis C, Achour L, Dupas H, Hatat C, and Rambaud JC (1993) Fate of
(1993) CD36 is a receptor for oxidized low density lipoprotein. J Biol Chem 268: beta-cyclodextrin in the human intestine. J Nutr 123:676–680.
11811–11816. Floyd AG (1999) Top ten considerations in the development of parenteral emulsions.
Engel GL, Farid NA, Faul MM, Richardson LA, and Winneroski LL (2000) Salt form Pharm Sci Technol Today 4:134–143.
selection and characterization of LY333531 mesylate monohydrate. Int J Pharm Flynn GL (1980) Buffers: pH control within pharmaceutical systems. J Parenter Drug
198:239–247. Assoc 34:139–162.
Erlich L, Yu D, Pallister DA, Levinson RS, Gole DG, Wilkinson PA, Erlich RE, Reeve LE, Flynn GL, Shah Y, Prakongpan S, Kwan KH, Higuchi WI, and Hofmann AF (1979)
and Viegas TX (1999) Relative bioavailability of danazol in dogs from liquid-filled Cholesterol solubility in organic solvents. J Pharm Sci 68:1090–1097.
hard gelatin capsules. Int J Pharm 179:49–53. Forbes RT, York P, and Davidson JR (1995) Dissolution kinetics and solubilities of
Ernst R and Arditti J (1980) Biological effects of surfactants. IV. Effects of non-ionics p-aminosalicylic acid and its salts. Int J Pharm 126:199–208.
and amphoterics on HeLa cells. Toxicology 15:233–242. Ford JL and Rubinstein MH (1979) Ageing of indomethacin-polyethylene glycol 6000
Esnaashari S, Javadzadeh Y, Batchelor HK, and Conway BR (2005) The use of solid dispersion. Pharm Acta Helv 54:353–358.
microviscometry to study polymer dissolution from solid dispersion drug delivery Forester GP, Tall AR, Bisgaier CL, and Glickman RM (1983) Rat intestine secretes
systems. Int J Pharm 292:227–230. spherical high density lipoproteins. J Biol Chem 258:5938–5943.
Essayan DM, Kagey-Sobotka A, Colarusso PJ, Lichtenstein LM, Ozols RF, and King ED Forster A, Hempenstall J, Tucker I, and Rades T (2001a) The potential of small-scale
(1996) Successful parenteral desensitization to paclitaxel. J Allergy Clin Immunol fusion experiments and the Gordon-Taylor equation to predict the suitability of
97:42–46. drug/polymer blends for melt extrusion. Drug Dev Ind Pharm 27:549–560.
Etter MC and Reutzel SM (1991) Hydrogen-bond directed cocrystallization and mo- Forster A, Hempenstall J, Tucker I, and Rades T (2001b) Selection of excipients for
lecular recognition properties of acyclic imides. J Am Chem Soc 113:2586–2598. melt extrusion with two poorly water-soluble drugs by solubility parameter cal-
Evans CH, Partyka M, and Van Stam J (2000) Naphthalene complexation by beta- culation and thermal analysis. Int J Pharm 226:147–161.
cyclodextrin: influence of added short chain branched and linear alcohols. J Incl Fort FL, Heyman IA, and Kesterson JW (1984) Hemolysis study of aqueous poly-
Phenom Macrocycl Chem 38:381–396. ethylene glycol 400, propylene glycol and ethanol combinations in vivo and in vitro.
Everett RM, Descotes G, Rollin M, Greener Y, Bradford JC, Benziger DP, and Ward SJ J Parenter Sci Technol 38:82–87.
(1993) Nephrotoxicity of pravadoline maleate (WIN 48098-6) in dogs: evidence of Frank DW, Gray JE, and Weaver RN (1976) Cyclodextrin nephrosis in the rat. Am J
maleic acid-induced acute tubular necrosis. Fundam Appl Toxicol 21:59–65. Pathol 83:367–382.
Fahelelbom KMS, Timoney RF, and Corrigan OI (1993) Micellar solubilization of Frank HS and Evans MW (1945) Free volume and entropy in condensed systems. 3.
clofazimine analogues in aqueous solutions of ionic and nonionic surfactants. Entropy in binary liquid mixtures : partial molal entropy in dilute dolutions:
Pharm Res 10:631–634. structure and thermodynamics in aqueous electrolytes. J Chem Phys 13:507–532.
Falchuk KH, Peterson L, and McNeil BJ (1985) Microparticulate-induced phlebitis: Freundlich H (1909) Kapillarchemie. p 144, Akademische Yerlagsgesellschaft,
its prevention by in-line filtration. N Engl J Med 312:78–82. Leipzig, Germany.
Fatouros DG, Bergenstahl B, and Mullertz A (2007a) Morphological observations on Friche E, Jensen PB, Sehested M, Demant EJ, and Nissen NN (1990a) The solvents
a lipid-based drug delivery system during in vitro digestion. Eur J Pharm Sci 31: cremophor EL and Tween 80 modulate daunorubicin resistance in the multidrug
85–94. resistant Ehrlich ascites tumor. Cancer Commun 2:297–303.
Fatouros DG, Deen GR, Arleth L, Bergenstahl B, Nielsen FS, Pedersen JS, Friche E, Jensen PB, Sehested M, Demant EJF, and Nissen NN (1990b) The solvents
and Mullertz A (2007b) Structural development of self nano emulsifying drug de- cremophor EL and Tween 80 modulate daunorubicin resistance in the multidrug
livery systems (SNEDDS) during in vitro lipid digestion monitored by small-angle resistant Ehrlich ascites tumor. Cancer Commun 2:297–303.
X-ray scattering. Pharm Res 24:1844–1853. Fridrikh SV, Yu JH, Brenner MP, and Rutledge GC (2003) Controlling the fiber
Fatouros DG and Mullertz A (2008) In vitro lipid digestion models in design of drug diameter during electrospinning. Phys Rev Lett 90:144502.
delivery systems for enhancing oral bioavailability. Expert Opin Drug Metab Friesen DT, Shanker R, Crew M, Smithey DT, Curatolo WJ, and Nightingale JAS
Toxicol 4:65–76. (2008) Hydroxypropyl methylcellulose acetate succinate-based spray-dried dis-
Fatouros DG, Nielsen FS, Douroumis D, Hadjileontiadis LJ, and Mullertz A (2008) In persions: an overview. Mol Pharm 5:1003–1019.
vitro-in vivo correlations of self-emulsifying drug delivery systems combining the Frijlink HW, Franssen EJF, Eissens AC, Oosting R, Lerk CF, and Meijer DKF (1991)
dynamic lipolysis model and neuro-fuzzy networks. Eur J Pharm Biopharm 69: The effects of cyclodextrins on the disposition of intravenously injected drugs in the
887–898. rat. Pharm Res 8:380–384.
Faucci MT, Melani F, and Mura P (2000) 1H-NMR and molecular modelling tech- Friscic T and Jones W (2010) Benefits of cocrystallisation in pharmaceutical mate-
niques for the investigation of the inclusion complex of econazole with alpha- rials science: an update. J Pharm Pharmacol 62:1547–1559.
cyclodextrin in the presence of malic acid. J Pharm Biomed Anal 23:25–31. Fu RC, Lidgate DM, Whatley JL, and McCullough T (1987) The biocompatibility of
Favvas EP and Mitropoulos AC (2008) What is spinodal decomposition? J Eng Sci parenteral vehicles: in vitro/in vivo screening comparison and the effect of excip-
Technol Rev 1:25–27. ients on hemolysis. J Parenter Sci Technol 41:164–168.
Febbraio M, Hajjar DP, and Silverstein RL (2001) CD36: a class B scavenger receptor Fukuda M, Miller DA, Peppas NA, and McGinity JW (2008) Influence of sulfobutyl
involved in angiogenesis, atherosclerosis, inflammation, and lipid metabolism. ether beta-cyclodextrin (Captisol) on the dissolution properties of a poorly soluble
J Clin Invest 108:785–791. drug from extrudates prepared by hot-melt extrusion. Int J Pharm 350:188–196.
Fedors RF (1974) Method for estimating both solubility parameters and molar vol- Fusaro F, Hanchen M, Mazzotti M, Muhrer G, and Subramaniam B (2005) Dense gas
umes of liquids. Polym Eng Sci 14:147–154. antisolvent precipitation: a comparative investigation of the GAS and PCA tech-
Fenyvesi F, Fenyvesi E, Szente L, Goda K, Bacso Z, Bacskay I, Varadi J, Kiss T, niques. Ind Eng Chem Res 44:1502–1509.
Molnar E, and Janaky T, et al. (2008) P-glycoprotein inhibition by membrane Gad SC, Cassidy CD, Aubert N, Spainhour B, and Robbe H (2006) Nonclinical vehicle
cholesterol modulation. Eur J Pharm Sci 34:236–242. use in studies by multiple routes in multiple species. Int J Toxicol 25:499–521.
Ferguson TFM and Prottey C (1976) The effect of surfactants upon mammalian cells Galcera J and Molins E (2009) Effect of the counterion on the solubility of iso-
in vitro. Food Cosmet Toxicol 14:431–434. structural pharmaceutical lamotrigine salts. Cryst Growth Des 9:327–334.
482 Williams et al.

Gallo-Torres HE, Ludorf J, and Brin M (1978) The effect of medium-chain trigly- Good DJ and Rodriguez-Hornedo N (2009) Solubility advantage of pharmaceutical
cerides on the bioavailability of vitamin E. Int J Vitam Nutr Res 48:240–241. cocrystals. Cryst Growth Des 9:2252–2264.
Ganta S, Paxton JW, Baguley BC, and Garg S (2009) Formulation and pharmaco- Goole J, Lindley DJ, Roth W, Carl SM, Amighi K, Kauffmann JM, and Knipp GT
kinetic evaluation of an asulacrine nanocrystalline suspension for intravenous (2010) The effects of excipients on transporter mediated absorption. Int J Pharm
delivery. Int J Pharm 367:179–186. 393:17–31.
Gao F, Zhang Z, Bu H, Huang Y, Gao Z, Shen J, Zhao C, and Li Y (2011a) Nano- Gordon M and Taylor JS (1952) Ideal copolymers and the 2nd-order transitions of
emulsion improves the oral absorption of candesartan cilexetil in rats: Performance synthetic rubbers. 1. Non-crystalline copolymers. Journal of Applied Chemistry 2:
and mechanism. J Control Release 149:168–174. 493–500.
Gao L, Liu GY, Wang XQ, Liu F, Xu YF, and Ma J (2011b) Preparation of a chemi- Gordon MS and Chowhan ZT (1987) Effect of tablet solubility and hygroscopicity on
cally stable quercetin formulation using nanosuspension technology. Int J Pharm disintegrant efficiency in direct compression tablets in terms of dissolution.
404:231–237. J Pharm Sci 76:907–909.
Gao L, Zhang DR, and Chen MH (2008) Drug nanocrystals for the formulation of Gordon MS and Chowhan ZT (1990) The effect of aging on disintegrant efficiency in
poorly soluble drugs and its application as a potential drug delivery system. direct compression tablets with varied solubility and hygroscopicity, in terms of
J Nanopart Res 10:845–862. dissolution. Drug Dev Ind Pharm 16:437–447.
Gao P (2008) Amorphous pharmaceutical solids: characterization, stabilization, and Gould PL (1986) Salt selection for basic drugs. Int J Pharm 33:201–217.
development of marketable formulations of poorly soluble drugs with improved Gould S and Scott RC (2005) 2-Hydroxypropyl-beta-cyclodextrin (HP-beta-CD):
oral absorption. Mol Pharm 5:903–904. a toxicology review. Food Chem Toxicol 43:1451–1459.
Gao P, Akrami A, Alvarez F, Hu J, Li L, Ma C, and Surapaneni S (2009) Charac- Graeser KA, Patterson JE, Zeitler JA, Gordon KC, and Rades T (2009) Correlating
terization and optimization of AMG 517 supersaturatable self-emulsifying drug thermodynamic and kinetic parameters with amorphous stability. Eur J Pharm
delivery system (S-SEDDS) for improved oral absorption. J Pharm Sci 98:516–528. Sci 37:492–498.
Gao P, Guyton ME, Huang T, Bauer JM, Stefanski KJ, and Lu Q (2004) Enhanced Grant DJW and Higuchi T (1990) Solubility Behavior of Organic Compounds, Wiley-
oral bioavailability of a poorly water soluble drug PNU-91325 by supersaturatable Interscience, New York.
formulations. Drug Dev Ind Pharm 30:221–229. Grantscharova E and Gutzow I (1986) Vapor-pressure, solubility and affinity of
Gao P and Morozowich W (2006) Development of supersaturatable self-emulsifying undercooled melts and glasses. J Non-Cryst Solids 81:99–127.
drug delivery system formulations for improving the oral absorption of poorly Green MR, Manikhas GM, Orlov S, Afanasyev B, Makhson AM, Bhar P, and Hawkins
soluble drugs. Expert Opin Drug Deliv 3:97–110. MJ (2006) Abraxane, a novel Cremophor-free, albumin-bound particle form of
Gao Y, Li ZG, Sun M, Guo CY, Yu AH, Xi YW, Cui J, Lou HX, and Zhai GX (2011c) paclitaxel for the treatment of advanced non-small-cell lung cancer. Ann Oncol 17:
Preparation and characterization of intravenously injectable curcumin nano- 1263–1268.
suspension. Drug Deliv 18:131–142. Greenhalgh DJ, Williams AC, Timmins P, and York P (1999) Solubility parameters
Gao Y, Li ZG, Sun M, Gao Y, Li ZG, Sun M, Li HL, Guo CY, Cui J, Li AG, Cao FL, Xi YW, as predictors of miscibility in solid dispersions. J Pharm Sci 88:1182–1190.
and Lou HX, et al. (2010a) Preparation, characterization, pharmacokinetics, and Griffin WC (1949) Classification of surface-active agents by HLB. J Soc Cosmet Chem
tissue distribution of curcumin nanosuspension with TPGS as stabilizer. Drug Dev 1:311–326.
Ind Pharm 36:1225–1234. Gross TD, Schaab K, Ouellette M, Zook S, Reddy JP, Shurtleff A, Sacaan AI, Alebic-
Gao YA, Qian SA, and Zhang JJ (2010b) Physicochemical and pharmacokinetic Kolbah T, and Bozigian H (2007) An approach to early-phase salt selection: ap-
characterization of a spray-dried cefpodoxime proxetil nanosuspension. Chem plication to NBI-75043. Org Process Res Dev 11:365–377.
Pharm Bull (Tokyo) 58:912–917. Grove M, Mullertz A, Nielsen JL, and Pedersen GP (2006) Bioavailability of
Garren KW and Pyter RA (1990) Aqueous solubility properties of a dibasic peptide- seocalcitol II: development and characterisation of self-microemulsifying drug de-
like compound. Int J Pharm 63:167–172. livery systems (SMEDDS) for oral administration containing medium and long
Gassmann P, List M, Schweitzer A, and Sucker H (1994) Hydrosols: alternatives for chain triglycerides. Eur J Pharmaceutical Sci 28:233–242.
the parenteral application of poorly water-soluble drugs. Eur J Pharm Biopharm Grove M, Pedersen GP, Nielsen JL, and Müllertz A (2005) Bioavailability of
40:64–72. seocalcitol I. Relating solubility in biorelevant media with oral bioavailability in
Gebauer D and Coelfen H (2011) Prenucleation clusters and non-classical nucleation. rats: effect of medium and long chain triglycerides. J Pharm Sci 94:1830–1838.
Nano Today 6:564–584. Groves MJ and de Galindez DA (1976) The self-emulsifying action of mixed surfac-
Gebauer D, Völkel A, and Cölfen H (2008) Stable prenucleation calcium carbonate tants in oil. Acta Pharm Suec 13:361–372.
clusters. Science 322:1819–1822. Grunenberg A, Henck JO, and Siesler HW (1996) Theoretical derivation and practical
Gelderblom H, Verweij J, Nooter K, and Sparreboom A (2001) Cremophor EL: the application of energy temperature diagrams as an instrument in preformulation
drawbacks and advantages of vehicle selection for drug formulation. Eur J Cancer studies of polymorphic drug substances. Int J Pharm 129:147–158.
37:1590–1598. Grzybowska K, Paluch M, Grzybowski A, Wojnarowska Z, Hawelek L, Kolodziejczyk K,
Gershkovich P, Fanous J, Qadri B, Yacovan A, Amselem S, and Hoffman A (2009) and Ngai KL (2010) Molecular dynamics and physical stability of amorphous anti-
The role of molecular physicochemical properties and apolipoproteins in associa- inflammatory drug: celecoxib. J Phys Chem B 114:12792–12801.
tion of drugs with triglyceride-rich lipoproteins: in-silico prediction of uptake by Gu CH, Rao D, Gandhi RB, Hilden J, and Raghavan K (2005) Using a novel
chylomicrons. J Pharm Pharmacol 61:31–39. multicompartment dissolution system to predict the effect of gastric pH on the
Gershkovich P and Hoffman A (2005) Uptake of lipophilic drugs by plasma derived oral absorption of weak bases with poor intrinsic solubility. J Pharm Sci 94:
isolated chylomicrons: linear correlation with intestinal lymphatic bioavailability. 199–208.
Eur J Pharm Sci 26:394–404. Gu Y, Wang GJ, Sun JG, Jia YW, Wang W, Xu MJ, Lv T, Zheng YT, and Sai Y (2009)
Gershkovich P and Hoffman A (2007) Effect of a high-fat meal on absorption and Pharmacokinetic characterization of ginsenoside Rh2, an anticancer nutrient from
disposition of lipophilic compounds: the importance of degree of association with ginseng, in rats and dogs. Food Chem Toxicol 47:2257–2268.
triglyceride-rich lipoproteins. Eur J Pharm Sci 32:24–32. Guerrieri P, Rumondor ACF, Li T, and Taylor LS (2010) Analysis of relationships
Ghosh I, Snyder J, Vippagunta R, Alvine M, Vakil R, Tong WQ, and Vippagunta S between solid-state properties, counterion, and developability of pharmaceutical
(2011) Comparison of HPMC based polymers performance as carriers for manu- salts. AAPS PharmSciTech 11:1212–1222.
facture of solid dispersions using the melt extruder. Int J Pharm 419:12–19. Guerrieri P and Taylor LS (2009) Role of salt and excipient properties on dispro-
Gibaud S, Zirar SB, Mutzenhardt P, Fries I, and Astier A (2005) Melarsoprol- portionation in the solid-state. Pharm Res 26:2015–2026.
cyclodextrins inclusion complexes. Int J Pharm 306:107–121. Gupta MK, Goldman D, Bogner RH, and Tseng YC (2001) Enhanced drug dissolution
Gibson L (2007) Lipid-based excipients for oral drug delivery, in Oral Lipid-Based and bulk properties of solid dispersions granulated with a surface adsorbent.
Formulations: Enhancing the Bioavailability of Poorly Water Soluble Drugs (Hauss Pharm Dev Technol 6:563–572.
DJ ed) pp 1–31, Informa Healthcare, Inc., New York. Gupta MK, Tseng YC, Goldman D, and Bogner RH (2002) Hydrogen bonding with
Gilis PA, de Conde V, and Vandecruys R (1997), inventors; Janssen Pharmaceutica adsorbent during storage governs drug dissolution from solid-dispersion granules.
NV, assignee. Beads Having a Core Coated with an Antifungal and a Polymer, U.S. Pharm Res 19:1663–1672.
patent 5633015. 1997 May 27. Gupta MK, Vanwert A, and Bogner RH (2003) Formation of physically stable
Giron D and Grant DJW (2002) Evaluation of solid-state properties of salts, in amorphous drugs by milling with Neusilin. J Pharm Sci 92:536–551.
Handbook of Pharmaceutical Salts: Properties, Selection and Use (Stahl PH Gupta P, Kakumanu VK, and Bansal AK (2004) Stability and solubility of celecoxib-
and Wermuth GH eds) pp 40–81, Wiley-VCH, Weinheim, Germany. PVP amorphous dispersions: a molecular perspective. Pharm Res 21:1762–1769.
Goldberg AH, Gibaldi M, and Kanig JL (1965) Increasing dissolution rates and Gupta PK (1996) Non-crystalline solids: glasses and amorphous solids. J Non-Cryst
gastrointestinal absorption of drugs via solid solutions and eutectic mixtures. I. Solids 195:158–164.
Theoretical considerations and discussion of the literature. J Pharm Sci 54: Gupta SK, Manfro RC, Tomlanovich SJ, Gambertoglio JG, Garovoy MR, and Benet
1145–1148. LZ (1990) Effect of food on the pharmacokinetics of cyclosporine in healthy subjects
Goldberg AH, Gibaldi M, and Kanig JL (1966a) Increasing dissolution rates and following oral and intravenous administration. J Clin Pharmacol 30:643–653.
gastrointestinal absorption of drugs via solid solutions and eutectic mixtures. II. Gupta SK and Benet LZ (1990) High-fat meals increase the clearance of cyclosporine.
Experimental evaluation of a eutectic mixture: urea-acetaminophen system. Pharm Res 7:46–48.
J Pharm Sci 55:482–487. Guyot M, Fawaz F, Bildet J, Bonini F, and Lagueny AM (1995) Physicochemical
Goldberg AH, Gibaldi M, and Kanig JL (1966b) Increasing dissolution rates and gas- characterization and dissolution of norfloxacin/cyclodextrin inclusion-compounds
trointestinal absorption of drugs via solid solutons and eutectic mixtures. III. Ex- and peg solid dispersions. Int J Pharm 123:53–63.
perimental evaluation of griseofulvin-succinic acid solid solution. J Pharm Sci 55: Guzmán H, Tawa M, Zhang Z, Ratanabanangkoon P, Shaw P, Mustonen P, Gardner CR,
487–492. Chen H, Moreau JP, Almarsson O, et al. (2004) A “spring and parachute” approach
Goldberg AH, Gibaldi M, Kanig JL, and Mayersohn M (1966c) Increasing dissolution to designing solid celecoxib formulations having enhanced oral absorption (Ab-
rates and gastrointestinal absorption of drugs via solid solutions and eutectic stract). AAPS PharmSciTech T2189.
mixtures. IV. Chloramphenicol-urea system. J Pharm Sci 55:581–583. Guzmán HR, Tawa M, Zhang Z, Ratanabanangkoon P, Shaw P, Gardner CR, Chen H,
Good D, Miranda C, and Rodriguez-Hornedo N (2011) Dependence of cocrystal for- Moreau JP, Almarsson O, and Remenar JF (2007) Combined use of crystalline salt
mation and thermodynamic stability on moisture sorption by amorphous polymer. forms and precipitation inhibitors to improve oral absorption of celecoxib from solid
CrystEngComm 13:1181–1189. oral formulations. J Pharm Sci 96:2686–2702.
Strategies for Low Drug Solubility 483
Gwak HS, Choi JS, and Choi HK (2005) Enhanced bioavailability of piroxicam via Higuchi WI, Parker AP, and Hamlin WE (1965b) Dissolution kinetics of a weak acid, 1,1-
salt formation with ethanolamines. Int J Pharm 297:156–161. hexamethylene P-tolylsulfonylsemicarbazide, and its sodium salt. J Pharm Sci 54:8–11.
Hajratwala BR (1982) Particle size reduction by a hammer mill. I. Effect of output Higuchi WI, Nelson E, and Wagner JG (1964) Solubility and dissolution rates in
screen size, feed particle size, and mill speed. J Pharm Sci 71:188–190. reactive media. J Pharm Sci 53:333–335.
Haleblian J and McCrone W (1969) Pharmaceutical applications of polymorphism. Hilkens PHE, Pronk LC, Verweij J, Vecht CJ, van Putten WL, and van den Bent MJ
J Pharm Sci 58:911–929. (1997a) Peripheral neuropathy induced by combination chemotherapy of docetaxel
Hall DG and Pethica BA (1967) Thermodynamics of micelle formation, in Nonionic and cisplatin. Br J Cancer 75:417–422.
Surfactants (Schick JM ed) pp 516–557, Marcel Dekker, New York. Hilkens PHE, Verweij J, Vecht CJ, Stoter G, and van den Bent MJ (1997b) Clinical
Hamosh M and Scow RO (1973) Lingual lipase and its role in the digestion of dietary characteristics of severe peripheral neuropathy induced by docetaxel (Taxotere).
lipid. J Clin Invest 52:88–95. Ann Oncol 8:187–190.
Han HK and Choi HK (2007) Improved absorption of meloxicam via salt formation Hippalgaonkar K, Majumdar S, and Kansara V (2010) Injectable lipid emulsions:
with ethanolamines. Eur J Pharm Biopharm 65:99–103. opportunities and challenges. AAPS PharmSciTech 11:1526–1540.
Han HK, Lee BJ, and Lee HK (2011) Enhanced dissolution and bioavailability of Hirasawa N, Okamoto H, and Danjo K (1999) Lactose as a low molecular weight
biochanin A via the preparation of solid dispersion: in vitro and in vivo evaluation. carrier of solid dispersions for carbamazepine and ethenzamide. Chem Pharm Bull
Int J Pharm 415:89–94. (Tokyo) 47:417–420.
Han SF, Yao TT, Zhang XX, Gan L, Zhu C, Yu HZ, and Gan Y (2009) Lipid-based Hirayama F, Usami M, Kimura K, and Uekama K (1997) Crystallization and poly-
formulations to enhance oral bioavailability of the poorly water-soluble drug morphic transition behavior of chloramphenicol palmitate in 2-hydroxypropyl-beta-
anethol trithione: effects of lipid composition and formulation. Int J Pharm 379: cyclodextrin matrix. Eur J Pharm Sci 5:23–30.
18–24. Hirayama F, Zaoh K, Harata K, Saenger W, and Uekama K (2001) Transparent,
Hanafy A, Spahn-Langguth H, Vergnault G, Grenier P, Tubic Grozdanis M, Lenhardt T, adhesive film formation of per-O-valeryl-beta-cyclodextrin. Chem Lett 636–637.
and Langguth P (2007) Pharmacokinetic evaluation of oral fenofibrate nano- Hirsch CA, Messenger RJ, and Brannon JL (1978) Fenoprofen: drug form selection
suspensions and SLN in comparison to conventional suspensions of micronized drug. and preformulation stability studies. J Pharm Sci 67:231–236.
Adv Drug Deliv Rev 59:419–426. Hirunpanich V, Katagi J, Sethabouppha B, and Sato H (2006) Demonstration of
Hancock BC and Parks M (2000) What is the true solubility advantage for amorphous docosahexaenoic acid as a bioavailability enhancer for CYP3A substrates: in vitro
pharmaceuticals? Pharm Res 17:397–404. and in vivo evidence using cyclosporin in rats. Drug Metab Dispos 34:305–310.
Hancock BC, Shamblin SL, and Zografi G (1995) Molecular mobility of amorphous Hirunpanich V, Murakoso K, and Sato H (2008) Inhibitory effect of docosahexaenoic
pharmaceutical solids below their glass transition temperatures. Pharm Res 12: acid (DHA) on the intestinal metabolism of midazolam: in vitro and in vivo studies
799–806. in rats. Int J Pharm 351:133–143.
Hancock BC and Zografi G (1997) Characteristics and significance of the amorphous Hirunpanich V and Sato H (2006) Docosahexaenoic acid (DHA) inhibits saquinavir
state in pharmaceutical systems. J Pharm Sci 86:1–12. metabolism in-vitro and enhances its bioavailability in rats. J Pharm Pharmacol
Hanke U, May K, Rozehnal V, Nagel S, Siegmund W, and Weitschies W (2010) 58:651–658.
Commonly used nonionic surfactants interact differently with the human efflux Hirunpanich V, Sethabouppha B, and Sato H (2007) Inhibitory effects of saturated
transporters ABCB1 (p-glycoprotein) and ABCC2 (MRP2). Eur J Pharm Biopharm and polyunsaturated fatty acids on the cytochrome P450 3A activity in rat liver
76:260–268. microsomes. Biol Pharm Bull 30:1586–1588.
Hanna M and York P (1998), inventors; University of Bradford, West Yorkshire, UK, Hitzenberger G and Jaschek I (1974) Comparative studies on the absorption of am-
assignee. Method and apparatus for the formulation of particles. US Patent picillin trihydrate and potassium ampicillin. Int J Clin Pharmacol 9:114–119.
5851453. 1998 Dec 22. Ho HO, Su HL, Tsai TM, and Sheu MT (1996) The preparation and characterization of
Hansen T, Holm P, and Schultz K (2004) Process characteristics and compaction of solid dispersions on pellets using a fluidized-bed system. Int J Pharm 139:223–229.
spray-dried emulsions containing a drug dissolved in lipid. Int J Pharm 287: Hofmann AF and Borgström B (1964) The intraluminal phase of fat digestion in man:
55–66. the lipid content of the micellar and oil phases of intestinal content obtained during
Hardung H, Djuric D, and Ali S (2010) Combining HME & solubilization: Soluplus— fat digestion and absorption. J Clin Invest 43:247–257.
the solid solution. Drug Delivery Technol 10:20–27. Hofmeister F (1888) Zur Lehre von der Wirkung der Salze. Arch Exp Pathol Phar-
Hargrove JT, Maxson WS, and Wentz AC (1989) Absorption of oral progesterone is makol 247–260.
influenced by vehicle and particle size. Am J Obstet Gynecol 161:948–951. Holm R and Hoest J (2004) Successful in silico predicting of intestinal lymphatic
Harrison JC and Eftink MR (1982) Cyclodextrin adamantanecarboxylate inclusion transfer. Int J Pharm 272:189–193.
complexes: a model system for the hydrophobic effect. Biopolymers 21:1153–1166. Holm R, Madsen JC, Shi W, Larsen KL, Stade LW, and Westh P (2011) Thermody-
Hasa D, Voinovich D, Perissutti B, Grassi M, Bonifacio A, Sergo V, Cepek C, namics of complexation of tauro- and glyco-conjugated bile salts with two modified
Chierotti MR, Gobetto R, and Dall’Acqua S, et al. (2011) Enhanced oral bio- beta-cyclodextrins. J Incl Phenom Macrocycl Chem 69:201–211.
availability of vinpocetine through mechanochemical salt formation: physico- Holmberg I, Aksnes L, Berlin T, Lindbäck B, Zemgals J, and Lindeke B (1990) Ab-
chemical characterization and in vivo studies. Pharm Res 28:1870–1883. sorption of a pharmacological dose of vitamin D3 from two different lipid vehicles
Hasegawa A, Nakagawa H, and Sugimoto I (1985) Bioavailability and stability of in man: comparison of peanut oil and a medium chain triglyceride. Biopharm Drug
nifedipine-enteric coating agent solid dispersion. Chem Pharm Bull (Tokyo) 33: Dispos 11:807–815.
388–391. Hong JY, Shah JC, and Mcgonagle MD (2011) Effect of cyclodextrin derivation and
Hastewell J, Lynch S, Fox R, Williamson I, Skeltonstroud P, and Mackay M (1994) amorphous state of complex on accelerated degradation of ziprasidone. J Pharm
Enhancement of human calcitonin absorption across the rat colon in vivo. Int J Sci 100:2703–2716.
Pharm 101:115–120. Hsieh Y-L, Ilevbare GA, Van Eerdenbrugh B, Box KJ, Sanchez-Felix MV, and Taylor LS
Haynes DA, Chisholm JA, Jones W, and Motherwell WDS (2004) Supramolecular (2012) pH-induced precipitation behavior of weakly basic compounds: determination of
synthon competition in organic sulfonates: a CSD survey. Cryst Eng Comm 6: extent and duration of supersaturation using potentiometric titration and correlation
584–588. to solid state properties. Pharm Res 29:2738–2753.
He Y, Li P, and Yalkowsky SH (2003) Solubilization of fluasterone in cosolvent/ Hu GX, Hu LF, Yang DZ, Li JW, Chen GR, Chen BB, Mruk DD, Bonanomi M,
cyclodextrin combinations. Int J Pharm 264:25–34. Silvestrini B, Cheng CY, et al. (2009) Adjudin targeting rabbit germ cell adhesion
He Y, Tabibi SE, and Yalkowsky SH (2006) Solubilization of two structurally related as a male contraceptive: a pharmacokinetics study. J Androl 30:87–93.
anticancer drugs: XK-469 and PPA. J Pharm Sci 95:97–107. Hu JH, Johnston KP, and Williams RO 3rd (2004a) Nanoparticle engineering pro-
Hecq J, Deleers M, Fanara D, Vranckx H, and Amighi K (2005) Preparation and cesses for enhancing the dissolution rates of poorly water soluble drugs. Drug Dev
characterization of nanocrystals for solubility and dissolution rate enhancement of Ind Pharm 30:233–245.
nifedipine. Int J Pharm 299:167–177. Hu JH, Johnston KP, and Williams RO 3rd (2004b) Rapid dissolving high potency danazol
Hecq J, Deleers M, Fanara D, Vranckx H, Boulanger P, Le Lamer S, and Amighi K powders produced by spray freezing into liquid process. Int J Pharm 271:145–154.
(2006) Preparation and in vitro/in vivo evaluation of nano-sized crystals for dis- Hu LD, Tang X, and Cui FD (2004c) Solid lipid nanoparticles (SLNs) to improve oral
solution rate enhancement of ucb-35440-3, a highly dosed poorly water-soluble bioavailability of poorly soluble drugs. J Pharm Pharmacol 56:1527–1535.
weak base. Eur J Pharm Biopharm 64:360–368. Hu Z, Prasad Yv R, Tawa R, Konishi T, Ishida M, Shibata N, and Takada K (2002)
Heerklotz H (2008) Interactions of surfactants with lipid membranes. Q Rev Biophys Diethyl ether fraction of Labrasol having a stronger absorption enhancing effect on
41:205–264. gentamicin than Labrasol itself. Int J Pharm 234:223–235.
Henck JO and Byrn SR (2007) Designing a molecular delivery system within a pre- Huang JJ, Li Y, Wigent RJ, Malick WA, Sandhu HK, Singhal D, and Shah NH (2011)
clinical timeframe. Drug Discov Today 12:189–199. Interplay of formulation and process methodology on the extent of nifedipine mo-
Heng D, Lee SH, Ng WK, and Tan RBH (2011) The nano spray dryer B-90. Expert lecular dispersion in polymers. Int J Pharm 420:59–67.
Opin Drug Deliv 8:965–972. Huang LF and Tong WQ (2004) Impact of solid state properties on developability
Hernell O, Staggers JE, and Carey MC (1990) Physical-chemical behavior of dietary assessment of drug candidates. Adv Drug Deliv Rev 56:321–334.
and biliary lipids during intestinal digestion and absorption. 2. Phase analysis and Huang N and Rodriguez-Hornedo N (2010) Effect of micellar solubilization on
aggregation states of luminal lipids during duodenal fat digestion in healthy adult cocrystal solubility and stability. Cryst Growth Des 10:2050–2053.
human beings. Biochemistry 29:2041–2056. Hugger ED, Novak BL, Burton PS, Audus KL, and Borchardt RT (2002) A compar-
Higashiyama M, Inada K, Ohtori A, and Tojo K (2004) Improvement of the ocular ison of commonly used polyethoxylated pharmaceutical excipients on their ability
bioavailability of timolol by sorbic acid. Int J Pharm 272:91–98. to inhibit P-glycoprotein activity in vitro. J Pharm Sci 91:1991–2002.
Higuchi T and Connons KA (1964) Advances in analytical chemistry and in- Hughey JR, DiNunzio JC, Bennett RC, Brough C, Miller DA, Ma H, Williams RO
strumentation, in Advances in Analytical Chemistry and Instrumentation (Reilly 3rd, and McGinity JW (2010) Dissolution enhancement of a drug exhibiting
CN ed) pp 117–212, Wiley-Interscience, New York. thermal and acidic decomposition characteristics by fusion processing: a com-
Higuchi WI and Hiestand EN (1963) Dissolution rates of finely divided drug powders. parative study of hot melt extrusion and KinetiSol dispersing. AAPS PharmS-
I. Effect of a distribution of particle sizes in a diffusion-controlled process. J Pharm ciTech 11:760–774.
Sci 52:67–71. Hülsmann S, Backensfeld T, Keitel S, and Bodmeier R (2000) Melt extrusion: an
Higuchi WI, Mir NA, and Desai SJ (1965a) Dissolution rates of polyphase mixtures. alternative method for enhancing the dissolution rate of 17beta-estradiol hemi-
J Pharm Sci 54:1405–1410. hydrate. Eur J Pharm Biopharm 49:237–242.
484 Williams et al.

Humberstone AJ, Porter CJH, and Charman WN (1996) A physicochemical basis for Jamil H, Dickson JK Jr, Chu CH, Lago MW, Rinehart JK, Biller SA, Gregg RE,
the effect of food on the absolute oral bioavailability of halofantrine. J Pharm Sci and Wetterau JR (1995) Microsomal triglyceride transfer protein: specificity of
85:525–529. lipid binding and transport. J Biol Chem 270:6549–6554.
Humberstone AJ, Porter CJH, Edwards GA, and Charman WN (1998) Association of Jang DJ, Jeong EJ, Lee HM, Kim BC, Lim SJ, and Kim CK (2006) Improvement of
halofantrine with postprandially derived plasma lipoproteins decreases its clear- bioavailability and photostability of amlodipine using redispersible dry emulsion.
ance relative to administration in the fasted state. J Pharm Sci 87:936–942. Eur J Pharm Sci 28:405–411.
Humphreys KJ and Rhodes CT (1968) Effect of temperature upon solubilization by Janknegt R, de Marie S, Bakker-Woudenberg IA, and Crommelin DJ (1992) Lipo-
a series of nonionic surfactants. J Pharm Sci 57:79–83. somal and lipid formulations of amphotericin B: clinical pharmacokinetics. Clin
Hurter PN, Scheutjens J, and Hatton TA (1993) Molecular modeling of micelle for- Pharmacokinet 23:279–291.
mation and solubilization in block-copolymer micelles. 1. A self-consistent mean- Jannin V, Musakhanian J, and Marchaud D (2008) Approaches for the development
field lattice theory. Macromolecules 26:5592–5601. of solid and semi-solid lipid-based formulations. Adv Drug Deliv Rev 60:734–746.
Hussain MM, Iqbal J, Anwar K, Rava P, and Dai K (2003) Microsomal triglyceride Janssens S, De Zeure A, Paudel A, Van Humbeeck J, Rombaut P, and Van den
transfer protein: a multifunctional protein. Front Biosci 8:s500–s506. Mooter G (2010) Influence of preparation methods on solid state supersaturation of
Hussain MM, Rava P, Pan X, Dai K, Dougan SK, Iqbal J, Lazare F, and Khatun I amorphous solid dispersions: a case study with itraconazole and Eudragit e100.
(2008) Microsomal triglyceride transfer protein in plasma and cellular lipid me- Pharm Res 27:775–785.
tabolism. Curr Opin Lipidol 19:277–284. Janssens S, Nagels S, Armas HN, D’Autry W, Van Schepdael A, and Van den Mooter G
Iacocca RG, Burcham CL, and Hilden LR (2010) Particle engineering: a strategy for (2008) Formulation and characterization of ternary solid dispersions made up of
establishing drug substance physical property specifications during small molecule Itraconazole and two excipients, TPGS 1000 and PVPVA 64, that were selected
development. J Pharm Sci 99:51–75. based on a supersaturation screening study. Eur J Pharm Biopharm 69:158–166.
Iervolino M, Raghavan SL, and Hadgraft J (2000) Membrane penetration enhance- Janssens S and Van den Mooter G (2009) Review: physical chemistry of solid dis-
ment of ibuprofen using supersaturation. Int J Pharm 198:229–238. persions. J Pharm Pharmacol 61:1571–1586.
Ignatious F and Baldoni J (2001), inventors; Smithkline Beecham Corporation, as- Jantratid E and Dressman J (2009) Biorelevant dissolution media simulating the
signee. Electrospun pharmaceutical compositions. Patent WO0154667. proximal human gastrointestinal tract: an update. Dissolution Technologies 16:
Ignatious F and Sun L (2004), inventors; Smithkline Beecham Corporation, assignee. 21–25.
Electrospun amorphous pharmaceutical compositions, Patent WO 2004014304. Jantratid E, Janssen N, Chokshi H, Tang K, and Dressman JB (2008) Designing
Ignatious F, Sun LH, Lee CP, and Baldoni J (2010) Electrospun nanofibers in oral biorelevant dissolution tests for lipid formulations: case example—lipid suspension
drug delivery. Pharm Res 27:576–588. of RZ-50. Eur J Pharm Biopharm 69:776–785.
Ikeda K, Tomida H, and Yotsuyanagi T (1977) Micellar interaction of tetracycline Jarho P, Pate DW, Brenneisen R, and Järvinen T (1998) Hydroxypropyl-beta-
antibiotics. Chem Pharm Bull (Tokyo) 25:1067–1072. cyclodextrin and its combination with hydroxypropyl-methylcellulose increases
Ingels F, Beck B, Oth M, and Augustijns P (2004) Effect of simulated intestinal fluid aqueous solubility of delta9-tetrahydrocannabinol. Life Sci 63:PL381–PL384.
on drug permeability estimation across Caco-2 monolayers. Int J Pharm 274: Järvinen T, Järvinen K, Schwarting N, and Stella VJ (1995) Beta-cyclodextrin
221–232. derivatives, SBE4-beta-CD and HP-beta-CD, increase the oral bioavailability of
Ingels F, Deferme S, Destexhe E, Oth M, Van den Mooter G, and Augustijns P (2002) cinnarizine in beagle dogs. J Pharm Sci 84:295–299.
Simulated intestinal fluid as transport medium in the Caco-2 cell culture model. Jaworek A (2007) Micro- and nanoparticle production by electrospraying. Powder
Int J Pharm 232:183–192. Technol 176:18–35.
Inoue Y, Liu Y, Tong LH, Shen BJ, and Jin DS (1993) Thermodynamics of molecular Jenning V and Gohla SH (2001) Encapsulation of retinoids in solid lipid nano-
recognition by cyclodextrins. 2. Calorimetric titration of inclusion complexation particles (SLN). J Microencapsul 18:149–158.
with modified beta-cyclodextrins: enthalpy-entropy compensation in host-guest Jia LJ, Zhang DR, Li ZY, Duan CX, Wang YC, Feng FF, Wang FH, Liu Y, and Zhang QA
complexation: from ionophore to cyclodextrin and cyclophane. J Am Chem Soc 115: (2010) Nanostructured lipid carriers for parenteral delivery of silybin: biodistribution
10637–10644. and pharmacokinetic studies. Colloids Surf B Biointerfaces 80:213–218.
Iqbal J, Anwar K, and Hussain MM (2003) Multiple, independently regulated Jia ZR, Lin P, Xiang Y, Wang XQ, Wang JC, Zhang X, and Zhang Q (2011) A novel
pathways of cholesterol transport across the intestinal epithelial cells. J Biol Chem nanomatrix system consisted of colloidal silica and pH-sensitive poly-
278:31610–31620. methylacrylate improves the oral bioavailability of fenofibrate. Eur J Pharm Bio-
Iqbal J and Hussain MM (2005) Evidence for multiple complementary pathways for pharm 79:126–134.
efficient cholesterol absorption in mice. J Lipid Res 46:1491–1501. Jinno J, Kamada N, Miyake M, Yamada K, Mukai T, Odomi M, Toguchi H, Liver-
Iqbal J and Hussain MM (2009) Intestinal lipid absorption. Am J Physiol Endocrinol sidge GG, Higaki K, and Kimura T (2006) Effect of particle size reduction on
Metab 296:E1183–E1194. dissolution and oral absorption of a poorly water-soluble drug, cilostazol, in beagle
Irie T and Uekama K (1997) Pharmaceutical applications of cyclodextrins. III. Tox- dogs. J Control Release 111:56–64.
icological issues and safety evaluation. J Pharm Sci 86:147–162. Jinno J, Kamada N, Miyake M, Yamada K, Mukai T, Odomi M, Toguchi H,
Ishikawa M and Hashimoto Y (2011) Improvement in aqueous solubility in small Liversidge GG, Higaki K, and Kimura T (2008) In vitro-in vivo correlation for wet-
molecule drug discovery programs by disruption of molecular planarity and sym- milled tablet of poorly water-soluble cilostazol. J Control Release 130:29–37.
metry. J Med Chem 54:1539–1554. Johari GP (2000) A resolution for the enigma of a liquid’s configurational entropy-
Ishikawa M, Yoshii H, and Furuta T (2005) Interaction of modified cyclodextrins with molecular kinetics relation. J Chem Phys 112:8958–8969.
cytochrome P-450. Biosci Biotechnol Biochem 69:246–248. Johari GP (2011a) Mechanical relaxation and the notion of time-dependent extent of
Islam MS and Narurkar MM (1993) Solubility, stability and ionization behaviour of ergodicity during the glass transition. Phys Rev E Stat Nonlin Soft Matter Phys 84:
famotidine. J Pharm Pharmacol 45:682–686. 021501.
Ismail AA, Gouda MW, and Motawi MM (1970) Micellar solubilization of barbiturates. Johari GP, Aji DPB, and Gunawan L (2011) Clausius limits on cooling and heating
I. Solubilities of certain barbiturates in polysorbates of varying hydrophobic chain through the liquid-glass range of three pharmaceuticals and one metal alloy:
length. J Pharm Sci 59:220–224. annealing effects and residual entropy. Thermochim Acta 522:173–181.
Ito Y, Kusawake T, Ishida M, Tawa R, Shibata N, and Takada K (2005) Oral solid Johari GP and Goldstein M (1970) Viscous liquids and glass transition. 2. Secondary
gentamicin preparation using emulsifier and adsorbent. J Control Release 105:23–31. relaxations in glasses of rigid molecules. J Chem Phys 53:2372–2388.
Ivanova R, Alexandridis P, and Lindman B (2001) Interaction of poloxamer block Johari GP, Kim S, and Shanker RM (2005) Dielectric studies of molecular motions in
copolymers with cosolvents and surfactants. Colloids and Surfaces a-Physicochem amorphous solid and ultraviscous acetaminophen. J Pharm Sci 94:2207–2223.
Eng Aspects 183:41–53. Johari GP and Shanker RM (2010) On determining the relaxation time of glass and
Jacobs C, Kayser O, and Müller RH (2000) Nanosuspensions as a new approach for amorphous pharmaceuticals’ stability from thermodynamic data. Thermochim
the formulation for the poorly soluble drug tarazepide. Int J Pharm 196:161–164. Acta 511:89–95.
Jagannath C, Wells A, Mshvildadze M, Olsen M, Sepulveda E, Emanuele M, Hunter Johnson BM, Charman WN, and Porter CJH (2002) An in vitro examination of the
RL Jr, and Dasgupta A (1999) Significantly improved oral uptake of amikacin in impact of polyethylene glycol 400, Pluronic P85, and vitamin E d-alpha-tocopheryl
FVB mice in the presence of CRL-1605 copolymer. Life Sci 64:1733–1738. polyethylene glycol 1000 succinate on P-glycoprotein efflux and enterocyte-based
Jain A, Ran YQ, and Yalkowsky SH (2004) Effect of pH-sodium lauryl sulfate com- metabolism in excised rat intestine. AAPS PharmSci 4:E40.
bination on solubilization of PG-300995 (an anti-HIV agent): a technical note. Johnson JLH, He Y, and Yalkowsky SH (2003) Prediction of precipitation-induced
AAPS PharmSciTech 5:e45. phlebitis: a statistical validation of an in vitro model. J Pharm Sci 92:1574–1581.
Jain A, Zheng W, Garad S, Weaver M, and Panicucci R (2010) Importance of early Johnson KC and Swindell AC (1996) Guidance in the setting of drug particle size
characterization of physicochemical properties in developing high-dose intravenous specifications to minimize variability in absorption. Pharm Res 13:1795–1798.
infusion regimens for poorly water-soluble compounds. PDA J Pharm Sci Technol Johnston JM and Rao GA (1965) Triglyceride biosynthesis in the intestinal mucosa.
64:517–526. Biochim Biophys Acta 106:1–9.
Jain N, Yang G, Tabibi SE, and Yalkowsky SH (2001) Solubilization of NSC-639829. Jones MN (1999) Surfactants in membrane solubilisation. Int J Pharm 177:137–159.
Int J Pharm 225:41–47. Jones PH, Rowley EK, Weiss AL, Bishop DL, and Chun AHC (1969) Insoluble
Jain NJ, Aswal VK, Goyal PS, and Bahadur P (2000) Salt induced micellization and erythromycin salts. J Pharm Sci 58:337–339.
micelle structures of PEO/PPO/PEO block copolymers in aqueous solution. Colloids Jones SP and Parr GD (1987) The acetotoluides as models for studying cyclodextrin
and Surfaces. a. Physicochemical and Engineering Aspects 173:85–94. inclusion complexes. Int J Pharm 36:223–231.
Jain NJ, George A, and Bahadur P (1999) Effect of salt on the micellization of Joshi HN, Tejwani RW, Davidovich M, Sahasrabudhe VP, Jemal M, Bathala MS,
pluronic P65 in aqueous solution. Colloids and Surfaces. A. Physicochemical and Varia SA, and Serajuddin ATM (2004) Bioavailability enhancement of a poorly
Engineering Aspects 157:275–283. water-soluble drug by solid dispersion in polyethylene glycol-polysorbate 80 mix-
Jambhekar S, Casella R, and Maher T (2004) The physicochemical characteristics ture. Int J Pharm 269:251–258.
and bioavailability of indomethacin from beta-cyclodextrin, hydroxyethyl-beta- Joshi MD and Müller RH (2009) Lipid nanoparticles for parenteral delivery of
cyclodextrin, and hydroxypropyl-beta-cyclodextrin complexes. Int J Pharm 270: actives. Eur J Pharm Biopharm 71:161–172.
149–166. Jouyban A, (2008). Review of the cosolvency models for predicting solubility of drugs
James PF (1985) Kinetics of crystal nucleation in silicate-glasses. J Non-Cryst Solids in water-cosolvent mixtures. Journal of Pharmacy and Pharmaceutical Sciences.
73:517–540. 11:32–57.
Strategies for Low Drug Solubility 485
Juhnke M, Berghausen J, and Timpe C (2010) Accelerated formulation development Kaushal AM and Bansal AK (2008) Thermodynamic behavior of glassy state of
for nanomilled active pharmaceutical ingredients using a screening approach. structurally related compounds. Eur J Pharm Biopharm 69:1067–1076.
Chem Eng Technol 33:1412–1418. Kauzmann W (1948) The nature of the glassy state and the behavior of liquids at low
Jun SW, Kim MS, Jo GH, Lee S, Woo JS, Park JS, and Hwang SJ (2005) Cefuroxime temperatures. Chem Rev 43:219–256.
axetil solid dispersions prepared using solution enhanced dispersion by super- Kawakami K (2009) Current status of amorphous formulation and other special
critical fluids. J Pharm Pharmacol 57:1529–1537. dosage forms as formulations for early clinical phases. J Pharm Sci 98:2875–2885.
Junco S, Casimiro T, Ribeiro N, Da Ponte MN, and Marques HMC (2002) Optimi- Kawakami K, Miyoshi K, and Ida Y (2004) Solubilization behavior of poorly soluble
sation of supercritical carbon dioxide systems for complexation of naproxen: beta- drugs with combined use of Gelucire 44/14 and cosolvent. J Pharm Sci 93:
cyclodextrin. J Incl Phenom Macrocycl Chem 44:69–73. 1471–1479.
Jung J and Perrut M (2001) Particle design using supercritical fluids: literature and Kawakami K, Oda N, Miyoshi K, Funaki T, and Ida Y (2006) Solubilization behavior
patent survey. J Supercrit Fluids 20:179–219. of a poorly soluble drug under combined use of surfactants and cosolvents. Eur J
Jung JY, Yoo SD, Lee SH, Kim KH, Yoon DS, and Lee KH (1999) Enhanced solubility Pharm Sci 28:7–14.
and dissolution rate of itraconazole by a solid dispersion technique. Int J Pharm Kayser O, Olbrich C, Yardley V, Kiderlen AF, and Croft SL (2003) Formulation of
187:209–218. amphotericin B as nanosuspension for oral administration. Int J Pharm 254:
Jung MS, Kim JS, Kim MS, Alhalaweh A, Cho W, Hwang SJ, and Velaga SP (2010) 73–75.
Bioavailability of indomethacin-saccharin cocrystals. J Pharm Pharmacol 62: Kearney AS, Mehta SC, and Radebaugh GW (1992) The effect of cyclodextrins on the
1560–1568. rate of intramolecular lactamization of gabapentin in aqueous-solution. Int J
Junghanns JU and Müller RH (2008) Nanocrystal technology, drug delivery and Pharm 78:25–34.
clinical applications. Int J Nanomedicine 3:295–309. Keck CM and Müller RH (2006) Drug nanocrystals of poorly soluble drugs produced
Junquera E, Tardajos G, and Aicart E (1993) Effect of the presence of beta- by high pressure homogenisation. Eur J Pharm Biopharm 62:3–16.
cyclodextrin on the micellization process of sodium dodecyl-sulfate or sodium Kemper EM, van Zandbergen AE, Cleypool C, Mos HA, Boogerd W, Beijnen JH,
perfluorooctanoate in water. Langmuir 9:1213–1219. and van Tellingen O (2003) Increased penetration of paclitaxel into the brain by
Kabanov AV, Batrakova EV, and Alakhov VY (2002) Pluronic block copolymers as inhibition of P-glycoprotein. Clin Cancer Res 9:2849–2855.
novel polymer therapeutics for drug and gene delivery. J Control Release 82: Kennedy M, Hu J, Gao PL, Li L, Ali-Reynolds A, Chal B, Gupta V, Ma C, Mahajan N,
189–212. and Akrami A, et al. (2008) Enhanced bioavailability of a poorly soluble VR1 an-
Kabanov AV, Batrakova EV, and Miller DW (2003) Pluronic block copolymers as tagonist using an amorphous solid dispersion approach: a case study. Mol Pharm 5:
modulators of drug efflux transporter activity in the blood-brain barrier. Adv Drug 981–993.
Deliv Rev 55:151–164. Kerc J and Srcic S (1995) Thermal-analysis of glassy pharmaceuticals. Thermochim
Kadam Y, Yerramilli U, and Bahadur A (2009) Solubilization of poorly water-soluble Acta 248:81–95.
drug carbamezapine in pluronic micelles: effect of molecular characteristics, tem- Kerc J, Srcic S, and Kofler B (1998) Alternative solvent-free preparation methods for
perature and added salt on the solubilizing capacity. Colloids Surf B Biointerfaces felodipine surface solid dispersions. Drug Dev Ind Pharm 24:359–363.
72:141–147. Kerns EH (2001) High throughput physicochemical profiling for drug discovery.
Kadam Y, Yerramilli U, Bahadur A, and Bahadur P (2011) Micelles from PEO-PPO- J Pharm Sci 90:1838–1858.
PEO block copolymers as nanocontainers for solubilization of a poorly water sol- Kerns EH and Di L (2003) Pharmaceutical profiling in drug discovery. Drug Discov
uble drug hydrochlorothiazide. Colloids Surf B Biointerfaces 83:49–57. Today 8:316–323.
Kamath AV, Chang M, Lee FY, Zhang YP, and Marathe PH (2005) Preclinical Keserü GM and Makara GM (2009) The influence of lead discovery strategies on the
pharmacokinetics and oral bioavailability of BMS-310705, a novel epothilone B properties of drug candidates. Nat Rev Drug Discov 8:203–212.
analog. Cancer Chemother Pharmacol 56:145–153. Kesisoglou F, Panmai S, and Wu YH (2007) Nanosizing: oral formulation de-
Kaminski K, Adrjanowicz K, Wojnarowska Z, Dulski M, Wrzalik R, Paluch M, velopment and biopharmaceutical evaluation. Adv Drug Deliv Rev 59:631–644.
Kaminska E, and Kasprzycka A (2011) Do intermolecular interactions control Kesisoglou F and Wu YH (2008) Understanding the effect of API properties on bio-
crystallization abilities of glass-forming liquids? J Phys Chem B 115:11537–11547. availability through absorption modeling. AAPS J 10:516–525.
Kanaujia P, Lau G, Ng WK, Widjaja E, Hanefeld A, Fischbach M, Maio M, and Tan Khalafallah N, Khalil SA, and Moustafa MA (1974) Bioavailability determination of
RBH (2011) Nanoparticle formation and growth during in vitro dissolution of two crystal forms of sulfameter in humans from urinary excretion data. J Pharm
ketoconazole solid dispersion. J Pharm Sci 100:2876–2885. Sci 63:861–864.
Kang JH, Oh DH, Oh Y-K, Yong CS, and Choi H-G (2012) Effects of solid carriers on Khan S, Batchelor H, Hanson P, Perrie Y, and Mohammed AR (2011) Physico-
the crystalline properties, dissolution SNEDDS). Eur J Pharm Biopharm 80: chemical characterisation, drug polymer dissolution and in vitro evaluation of
289–297. phenacetin and phenylbutazone solid dispersions with polyethylene glycol 8000.
Kapetanovic IM, Muzzio M, Hu SC, Crowell JA, Rajewski RA, Haslam JL, Jong L, J Pharm Sci 100:4281–4294.
and McCormick DL (2010) Pharmacokinetics and enhanced bioavailability of Khoo SM, Edwards GA, Porter CJH, and Charman WN (2001) A conscious dog model
candidate cancer preventative agent, SR13668 in dogs and monkeys. Cancer for assessing the absorption, enterocyte-based metabolism, and intestinal lym-
Chemother Pharmacol 65:1109–1116. phatic transport of halofantrine. J Pharm Sci 90:1599–1607.
Kapsi SG and Ayres JW (2001) Processing factors in development of solid solution Khoo SM, Porter CJH, and Charman WN (2000) The formulation of halo-
formulation of itraconazole for enhancement of drug dissolution and bio- fantrine as either non-solubilizing PEG 6000 or solubilizing lipid based solid dis-
availability. Int J Pharm 229:193–203. persions: physical stability and absolute bioavailability assessment. Int J Pharm
Karavas E, Georgarakis E, Bikiaris D, Thomas T, Katsos V, and Xenakis A (2001) 205:65–78.
Hydrophilic matrices as carriers in felodipine solid dispersion systems, in Trends in Khoo SM, Shackleford DM, Porter CJH, Edwards GA, and Charman WN (2003)
Colloid and Interface Science XV (Koutsoukos PG ed) pp 149–152, Springer, Berlin. Intestinal lymphatic transport of halofantrine occurs after oral administration of
Karavas E, Georgarakis E, Sigalas MP, Avgoustakis K, and Bikiaris D (2007) a unit-dose lipid-based formulation to fasted dogs. Pharm Res 20:1460–1465.
Investigation of the release mechanism of a sparingly water-soluble drug from Khougaz K and Clas SD (2000) Crystallization inhibition in solid dispersions of MK-
solid dispersions in hydrophilic carriers based on physical state of drug, particle 0591 and poly(vinylpyrrolidone) polymers. J Pharm Sci 89:1325–1334.
size distribution and drug-polymer interactions. Eur J Pharm Biopharm 66: Ki MH, Choi MH, Ahn KB, Kim BS, Im DS, Ahn SK, and Shin HJ (2008) The efficacy
334–347. and safety of clopidogrel resinate as a novel polymeric salt form of clopidogrel. Arch
Karavas E, Ktistis G, Xenakis A, and Georgarakis E (2005) Miscibility behavior and Pharm Res 31:250–258.
formation mechanism of stabilized felodipine-polyvinylpyrrolidone amorphous Kim JH, Lee SK, Ki MH, Choi WK, Ahn SK, Shin HJ, and Hong CI (2004a) De-
solid dispersions. Drug Dev Ind Pharm 31:473–489. velopment of parenteral formulation for a novel angiogenesis inhibitor, CKD-732
Karpf DM, Holm R, Garafalo C, Levy E, Jacobsen J, and Müllertz A (2006) Effect of through complexation with hydroxypropyl-beta-cyclodextrin. Int J Pharm 272:
different surfactants in biorelevant medium on the secretion of a lipophilic com- 79–89.
pound in lipoproteins using Caco-2 cell culture. J Pharm Sci 95:45–55. Kim SC, Kim DW, Shim YH, Bang JS, Oh HS, Wan Kim S, and Seo MH (2001) In vivo
Kashchiev D and van Rosmalen GM (2003) Review: nucleation in solutions revisited. evaluation of polymeric micellar paclitaxel formulation: toxicity and efficacy.
Cryst Res Technol 38:555–574. J Control Release 72:191–202.
Kasim NA, Whitehouse M, Ramachandran C, Bermejo M, Lennernäs H, Hussain AS, Kim TY, Kim DW, Chung JY, Shin SG, Kim SC, Heo DS, Kim NK, and Bang YJ
Junginger HE, Stavchansky SA, Midha KK, and Shah VP, et al. (2004) Molecular (2004b) Phase I and pharmacokinetic study of Genexol-PM, a cremophor-free,
properties of WHO essential drugs and and bioavailability of flurbiprofen in solid polymeric micelle-formulated paclitaxel, in patients with advanced malignancies.
self-nanoemulsifying drug delivery system (solid provisional biopharmaceutical Clin Cancer Res 10:3708–3716.
classification. Mol Pharm 1:85–96. Kim Y, Oksanen DA, Massefski W Jr, Blake JF, Duffy EM, and Chrunyk B (1998)
Kataoka K, Harada A, and Nagasaki Y (2001) Block copolymer micelles for drug Inclusion complexation of ziprasidone mesylate with beta-cyclodextrin sulfobutyl
delivery: design, characterization and biological significance. Adv Drug Deliv Rev ether. J Pharm Sci 87:1560–1567.
47:113–131. Kimura F, Murakami S, and Fujishiro R (1975) Thermodynamics of aqueous-
Katneni K, Charman SA, and Porter CJH (2007) Impact of cremophor-EL and solutions of nonelectrolytes. 2. Enthalpies of transfer of 1-methyl-2-pyrrolidinone
polysorbate-80 on digoxin permeability across rat jejunum: delineation of ther- from water to many aqueous alcohols. J Solution Chem 4:241–247.
modynamic and transporter related events using the reciprocal permeability ap- Kimura K, Hirayama F, Arima H, and Uekama K (1999) Solid-state 13C nuclear
proach. J Pharm Sci 96:280–293. magnetic resonance spectroscopic study on amorphous solid complexes of tolbu-
Kato Y and Kohketsu M (1981) Relationship between polymorphism and bio- tamide with 2-hydroxypropyl-alpha- and -beta-cyclodextrins. Pharm Res 16:
availability of amorbarbital in the rabbit. Chem Pharm Bull (Tokyo) 29:268–272. 1729–1734.
Kaukonen AM, Boyd BJ, Charman WN, and Porter CJH (2004a) Drug solubilization Kindel T, Lee DM, and Tso P (2010) The mechanism of the formation and secretion of
behavior during in vitro digestion of suspension formulations of poorly water- chylomicrons. Atheroscler Suppl 11:11–16.
soluble drugs in triglyceride lipids. Pharm Res 21:254–260. King SYP, Basista AM, and Torosian G (1989) Self-association and solubility
Kaukonen AM, Boyd BJ, Porter CJH, and Charman WN (2004b) Drug solubilization behaviors of a novel anticancer agent, brequinar sodium. J Pharm Sci 78:95–100.
behavior during in vitro digestion of simple triglyceride lipid solution formulations. Kipp JE (2004) The role of solid nanoparticle technology in the parenteral delivery of
Pharm Res 21:245–253. poorly water-soluble drugs. Int J Pharm 284:109–122.
486 Williams et al.

Kipp JE (2007) Solubilizing systems for parenteral formulation development - Ku S (2010) Salt and polymorph selection strategy based on the Biopharmaceutical
small molecules, in Solvent Systems and Their Selection in Pharmaceutics and Classification System for early pharmaceutical development. Am Pharmaceutical
Biopharmaceutics (Augustijns P and Brewster M eds) pp 309–336, Springer, Rev 20:30.
New York. Ku S and Velagaleti R (2010) Solutol HS15 as a novel excipient. Pharm Technol 108:
Kipp JE, Wong JCTW, and Doty MJ, and Rebbeck. CL (2002), inventors; Baxter 110.
International Inc., assignee. Microprecipitation method for preparing submicron Kumar L, Amin A, and Bansal AK (2007) An overview of automated systems relevant
suspensions U.S. patent 6607784. 2003 Aug 19. in pharmaceutical salt screening. Drug Discov Today 12:1046–1053.
Klein CE, Chiu YL, Awni W, Zhu T, Heuser RS, Doan T, Breitenbach J, Morris JB, Kumar L, Amin A, and Bansal AK (2008a) Preparation and characterization of salt
Brun SC, and Hanna GJ (2007) The tablet formulation of lopinavir/ritonavir pro- forms of enalapril. Pharm Dev Technol 13:345–357.
vides similar bioavailability to the soft-gelatin capsule formulation with less Kumar L, Amin A, and Bansal AK (2008b) Salt selection in drug development. Pharm
pharmacokinetic variability and diminished food effect. J Acquir Immune Defic Technol 3:128–146.
Syndr 44:401–410. Kumar NS and Mansbach CM 2nd (1999) Prechylomicron transport vesicle: isolation
Koester LS, Bertuol JB, Groch KR, Xavier CR, Moellerke R, Mayorga P, Dalla Costa and partial characterization. Am J Physiol 276:G378–G386.
T, and Bassani VL (2004) Bioavailability of carbamazepine:beta-cyclodextrin Kurkov SV, Ukhatskaya EV, and Loftsson T (2011) Drug/cyclodextrin: beyond in-
complex in beagle dogs from hydroxypropylmethylcellulose matrix tablets. Eur J clusion complexation. J Incl Phenom Macrocycl Chem 69:297–301.
Pharm Sci 22:201–207. Kwei GY, Novak LB, Hettrick LA, Reiss ER, Ostovic D, Loper AE, Lui CY, Higgins
Koester LS, Xavier CR, Mayorga P, and Bassani VL (2003) Influence of beta- RJ, Chen IW, and Lin JH (1995) Regiospecific intestinal absorption of the HIV
cyclodextrin complexation on carbamazepine release from hydroxypropyl methyl- protease inhibitor L-735,524 in beagle dogs. Pharm Res 12:884–888.
cellulose matrix tablets. Eur J Pharm Biopharm 55:85–91. Kwon GS and Forrest ML (2006) Amphiphilic block copolymer micelles for nanoscale
Kohan A, Yoder S, and Tso P (2010) Lymphatics in intestinal transport of nutrients drug delivery. Drug Dev Res 67:15–22.
and gastrointestinal hormones. Ann N Y Acad Sci 1207 (Suppl 1):E44–E51. Ladman AJ, Padykula HA, and Strauss EW (1963) A morphological study of fat
Kojima T, Tsutsumi S, Yamamoto K, Ikeda Y, and Moriwaki T (2010) High- transport in the normal human jejunum. Am J Anat 112:389–419.
throughput cocrystal slurry screening by use of in situ Raman microscopy and Lai F, Sinico C, Ennas G, Marongiu F, Marongiu G, and Fadda AM (2009) Diclofenac
multi-well plate. Int J Pharm 399:52–59. nanosuspensions: influence of preparation procedure and crystal form on drug
Koltun M, Morizzi J, Katneni K, Charman SA, Shackleford DM, and McIntosh MP dissolution behaviour. Int J Pharm 373:124–132.
(2010) Preclinical comparison of intravenous melphalan pharmacokinetics ad- Lakshman JP, Cao Y, Kowalski J, and Serajuddin ATM (2008) Application of melt
ministered in formulations containing either (SBE)7cm-b-cyclodextrin or a co- extrusion in the development of a physically and chemically stable high-energy
solvent system. Biopharm Drug Dispos 31:450–454. amorphous solid dispersion of a poorly water-soluble drug. Mol Pharm 5:994–1002.
Komiyama M and Bender ML (1978) Importance of apolar binding in complex- Langguth P, Hanafy A, Frenzel D, Grenier P, Nhamias A, Ohlig T, Vergnault G,
formation of cyclodextrins with adamantanecarboxylate. J Am Chem Soc 100: and Spahn-Langguth H (2005) Nanosuspension formulations for low-soluble drugs:
2259–2260. pharmacokinetic evaluation using spironolactone as model compound. Drug Dev
Kondo N, Iwao T, Masuda H, Yamanouchi K, Ishihara Y, Yamada N, Haga T, Ogawa Ind Pharm 31:319–329.
Y, and Yokoyama K (1993) Improved oral absorption of a poorly water-soluble Larsen A, Holm R, Pedersen ML, and Müllertz A (2008a) Lipid-based formulations
drug, HO-221, by wet-bead milling producing particles in submicron region. Chem for danazol containing a digestible surfactant, Labrafil M2125CS: in vivo bio-
Pharm Bull (Tokyo) 41:737–740. availability and dynamic in vitro lipolysis. Pharm Res 25:2769–2777.
Konno H, Handa T, Alonzo DE, and Taylor LS (2008) Effect of polymer type on the Larsen AT, Sassene P, and Müllertz A (2011) In vitro lipolysis models as a tool for the
dissolution profile of amorphous solid dispersions containing felodipine. Eur J characterization of oral lipid and surfactant based drug delivery systems. Int J
Pharm Biopharm 70:493–499. Pharm 417:245–255.
Konno H and Taylor LS (2006) Influence of different polymers on the crystallization Larsen C, Larsen SW, Jensen H, Yaghmur A, and Ostergaard J (2009) Role of in vitro
tendency of molecularly dispersed amorphous felodipine. J Pharm Sci 95: release models in formulation development and quality control of parenteral
2692–2705. depots. Expert Opin Drug Deliv 6:1283–1295.
Konno H and Taylor LS (2008) Ability of different polymers to inhibit the crystal- Larsen C, Ostergaard J, Larsen SW, Jensen H, Jacobsen S, Lindegaard C, and
lization of amorphous felodipine in the presence of moisture. Pharm Res 25: Andersen PH (2008b) Intra-articular depot formulation principles: role in the man-
969–978. agement of postoperative pain and arthritic disorders. J Pharm Sci 97:4622–4654.
Kossena GA, Boyd BJ, Porter CJH, and Charman WN (2003) Separation and char- Larsen SW, Østergaard J, Poulsen SV, Schulz B, and Larsen C (2007) Diflunisal salts
acterization of the colloidal phases produced on digestion of common formulation of bupivacaine, lidocaine and morphine: use of the common ion effect for prolonging
lipids and assessment of their impact on the apparent solubility of selected poorly the release of bupivacaine from mixed salt suspensions in an in vitro dialysis
water-soluble drugs. J Pharm Sci 92:634–648. model. Eur J Pharm Sci 31:172–179.
Kossena GA, Charman WN, Boyd BJ, Dunstan DE, and Porter CJH (2004) Probing Larson KA and King ML (1986) Evaluation of supercritical fluid extraction in the
drug solubilization patterns in the gastrointestinal tract after administration of pharmaceutical industry. Biotechnol Prog 2:73–82.
lipid-based delivery systems: a phase diagram approach. J Pharm Sci 93:332–348. Lasic DD (1996) Doxorubicin in sterically stabilized liposomes. Nature 380:561–562.
Kossena GA, Charman WN, Wilson CG, O’Mahony B, Lindsay B, Hempenstall JM, Lauer ME, Grassmann O, Siam M, Tardio J, Jacob L, Page S, Kindt JH, Engel A,
Davison CL, Crowley PJ, and Porter CJH (2007) Low dose lipid formulations: and Alsenz J (2011) Atomic force microscopy-based screening of drug-excipient
effects on gastric emptying and biliary secretion. Pharm Res 24:2084–2096. miscibility and stability of solid dispersions. Pharm Res 28:572–584.
Kostewicz ES, Wunderlich M, Brauns U, Becker R, Bock T, and Dressman JB (2004) Laughlin RG (1994) The Aqueous Phase Behaviour of Surfactants, Academic Press,
Predicting the precipitation of poorly soluble weak bases upon entry in the small London.
intestine. J Pharm Pharmacol 56:43–51. Law D, Krill SL, Schmitt EA, Fort JJ, Qiu YH, Wang WL, and Porter WR (2001)
Kou W, Cai CF, Xu SY, Wang H, Liu J, Yang D, and Zhang TH (2011) In vitro and in Physicochemical considerations in the preparation of amorphous ritonavir-poly
vivo evaluation of novel immediate release carbamazepine tablets: complexation (ethylene glycol) 8000 solid dispersions. J Pharm Sci 90:1015–1025.
with hydroxypropyl-b-cyclodextrin in the presence of HPMC. Int J Pharm 409: Law D, Schmitt EA, Marsh KC, Everitt EA, Wang WL, Fort JJ, Krill SL, and Qiu YH
75–80. (2004) Ritonavir-PEG 8000 amorphous solid dispersions: in vitro and in vivo
Kovács K, Stampf G, Klebovich I, Antal I, and Ludányi K (2009) Aqueous solvent evaluations. J Pharm Sci 93:563–570.
system for the solubilization of azole compounds. Eur J Pharm Sci 36:352–358. Law D, Wang WL, Schmitt EA, and Long MA (2002) Prediction of poly(ethylene)
Kovarik JM, Mueller EA, van Bree JB, Tetzloff W, and Kutz K (1994) Reduced inter- glycol-drug eutectic compositions using an index based on the van’t Hoff equation.
and intraindividual variability in cyclosporine pharmacokinetics from a micro- Pharm Res 19:315–321.
emulsion formulation. J Pharm Sci 83:444–446. Law SL, Lo WY, Lin FM, and Chaing CH (1992) Dissolution and absorption of ni-
Kozlov MY, Melik-Nubarov NS, Batrakova EV, and Kabanov AV (2000) Relationship fedipine in polyethylene-glycol solid dispersion containing phosphatidylcholine. Int
between pluronic block copolymer structure, critical micellization concentration J Pharm 84:161–166.
and partitioning coefficients of low molecular mass solutes. Macromolecules 33: Lawrence MJ (1994) Surfactant systems: their use in drug-delivery. Chem Soc Rev
3305–3313. 23:417–424.
Kramer SF and Flynn GL (1972) Solubility of organic hydrochlorides. J Pharm Sci Ledwidge MT and Corrigan OI (1998) Effects of surface active characteristics and
61:1896–1904. solid state forms on the pH solubility profiles of drug-salt systems. Int J Pharm
Krauel-Goellner K (2008) Strategies to advance the bioavailability of low solubility 174:187–200.
drugs: increasing solubility using nanocrystals, in Controlled Release Society An- Lee YHP, Sathigari S, Lin YJJ, Ravis WR, Chadha G, Parsons DL, Rangari VK,
nual Meeting, 2008 12–13 July, CRS Workshop, New York. Wright N, and Babu RJ (2009) Gefitinibcyclodextrin inclusion complexes: physico-
Krause KP and Müller RH (2001) Production and characterisation of highly con- chemical characterization and dissolution studies. Drug Dev Ind Pharm 35:
centrated nanosuspensions by high pressure homogenisation. Int J Pharm 214: 1113–1120.
21–24. Lee YC, Zocharski PD, and Samas B (2003) An intravenous formulation decision tree
Krieger M (2001) Scavenger receptor class B type I is a multiligand HDL receptor for discovery compound formulation development. Int J Pharm 253:111–119.
that influences diverse physiologic systems. J Clin Invest 108:793–797. Lehner R and Kuksis A (1996) Biosynthesis of triacylglycerols. Prog Lipid Res 35:
Krishna AK and Flanagan DR (1989) Micellar solubilization of a new antimalarial 169–201.
drug, beta-arteether. J Pharm Sci 78:574–576. Leonard GD, Fojo T, and Bates SE (2003) The role of ABC transporters in clinical
Kriwet K and Muller-Goymann CC (1993) Binary diclofenac diethylamine water- practice. Oncologist 8:411–424.
systems: micelles, vesicles and lyotropic liquid-crystals. Eur J Pharm Biopharm Lespine A, Chanoit G, Bousquet-Melou A, Lallemand E, Bassissi FM, Alvinerie M,
39:234–238. and Toutain PL (2006) Contribution of lymphatic transport to the systemic expo-
Krzyzaniak JF, Raymond DM, and Yalkowsky SH (1997) Lysis of human red blood sure of orally administered moxidectin in conscious lymph duct-cannulated dogs.
cells. 2. Effect of contact time on cosolvent induced hemolysis. Int J Pharm 152: Eur J Pharm Sci 27:37–43.
193–200. Leucht C, Heres S, Kane JM, Kissling W, Davis JM, and Leucht S (2011) Oral versus
Krzyzaniak JF and Yalkowsky SH (1998) Lysis of human red blood cells. 3: Effect of depot antipsychotic drugs for schizophrenia: a critical systematic review and meta-
contact time on surfactant-induced hemolysis. PDA J Pharm Sci Technol 52:66–69. analysis of randomised long-term trials. Schizophr Res 127:83–92.
Strategies for Low Drug Solubility 487
Leuner C and Dressman J (2000) Improving drug solubility for oral delivery using Lindfors L, Forssén S, Westergren J, and Olsson U (2008) Nucleation and crystal
solid dispersions. Eur J Pharm Biopharm 50:47–60. growth in supersaturated solutions of a model drug. J Colloid Interface Sci 325:
Li-Blatter X, Nervi P, and Seelig A (2009) Detergents as intrinsic P-glycoprotein 404–413.
substrates and inhibitors. Biochim Biophys Acta 1788:2335–2344. Lindman B, Wennerstrom H, Gustavsson H, Kamenka N, and Brun B (1980)
Li BQ, Dong X, Fang SH, Gao JY, Yang GQ, and Zhao H (2011a) Systemic toxicity Some aspects on the hydration of surfactant micelles. Pure Appl Chem 52:
and toxicokinetics of a high dose of polyethylene glycol 400 in dogs following in- 1307–1315.
travenous injection. Drug Chem Toxicol 34:208–212. Lindmark T, Kimura Y, and Artursson P (1998) Absorption enhancement through
Li C-C, Lii C-K, Liu K-L, Yang J-J, and Chen H-W (2006) n-6 and n-3 poly- intracellular regulation of tight junction permeability by medium chain fatty acids
unsaturated fatty acids down-regulate cytochrome P-450 2B1 gene expression in Caco-2 cells. J Pharmacol Exp Ther 284:362–369.
induced by phenobarbital in primary rat hepatocytes. J Nutr Biochem 17: Lindquist S and Hernell O (2010) Lipid digestion and absorption in early life: an
707–715. update. Curr Opin Clin Nutr Metab Care 13:314–320.
Li C-C, Lii C-K, Liu K-L, Yang J-J, and Chen H-W (2007) DHA down-regulates Linn M, Collnot EM, Djuric D, Hempel K, Fabian E, Kolter K, and Lehr CM (2012)
phenobarbital-induced cytochrome P450 2B1 gene expression in rat primary hep- Soluplus® as an effective absorption enhancer of poorly soluble drugs in vitro and
atocytes by attenuating CAR translocation. Toxicol Appl Pharmacol 225:329–336. in vivo. Eur J Pharm Sci 45:336–343.
Li HL, Zhao XB, Ma YK, Zhai GX, Li LB, and Lou HX (2009a) Enhancement of Lipinski CA (2000) Drug-like properties and the causes of poor solubility and poor
gastrointestinal absorption of quercetin by solid lipid nanoparticles. J Control permeability. J Pharmacol Toxicol Methods 44:235–249.
Release 133:238–244. Lipinski CA, Lombardo F, Dominy BW, and Feeney PJ (1997) Experimental and
Li P, Hynes SR, Haefele TF, Pudipeddi M, Royce AE, and Serajuddin ATM (2009b) computational approaches to estimate solubility and permeability in drug discov-
Development of clinical dosage forms for a poorly water-soluble drug II: formula- ery and development settings. Adv Drug Deliv Rev 23:3–25.
tion and characterization of a novel solid microemulsion preconcentrate system for List M and Sucker H (1988) inventors; Sandoz Ltd. Basel Switzerland, assignee.
oral delivery of a poorly water-soluble drug. J Pharm Sci 98:1750–1764. Pharmaceutical colloidal hydrosols for injection. GB Patent 2,200,048.
Li P, Patel H, Tabibi SE, Vishnuvajjala R, and Yalkowsky SH (1999a) Evaluation of Liu JS, Rigsbee DR, Stotz C, and Pikal MJ (2002) Dynamics of pharmaceutical
intravenous flavopiridol formulations. PDA J Pharm Sci Technol 53:137–140. amorphous solids: the study of enthalpy relaxation by isothermal microcalorime-
Li P, Tabibi SE, and Yalkowsky SH (1999b) Solubilization of ionized and un-ionized try. J Pharm Sci 91:1853–1862.
flavopiridol by ethanol and polysorbate 20. J Pharm Sci 88:507–509. Liu L and Guo QX (2002) The driving forces in the inclusion complexation of cyclo-
Li P, Tabibi SE, and Yalkowsky SH (1998a) Combined effect of complexation and pH dextrins. J Incl Phenom Macrocycl Chem 42:1–14.
on solubilization. J Pharm Sci 87:1535–1537. Liu P, Rong XY, Laru J, van Veen B, Kiesvaara J, Hirvonen J, Laaksonen T,
Li P, Tabibi SE, and Yalkowsky SH (1999c) Solubilization of flavopiridol by pH and Peltonen L (2011) Nanosuspensions of poorly soluble drugs: preparation and
control combined with cosolvents, surfactants, or complexants. J Pharm Sci 88: development by wet milling. Int J Pharm 411:215–222.
945–947. Liversidge GG and Cundy KC (1995) Particle-size reduction for improvement of oral
Li P, Vishnuvajjala R, Tabibi SE, and Yalkowsky SH (1998b) Evaluation of in vitro bioavailability of hydrophobic drugs. 1. Absolute oral bioavailability of nano-
precipitation methods. J Pharm Sci 87:196–199. crystalline danazol in beagle dogs. Int J Pharm 125:91–97.
Li P and Zhao L (2003) Solubilization of flurbiprofen in pH-surfactant solutions. Lloyd GR, Craig DQM, and Smith A (1997) An investigation into the melting be-
J Pharm Sci 92:951–956. havior of binary mixes and solid dispersions of paracetamol and PEG 4000.
Li P and Zhao L (2007) Developing early formulations: practice and perspective. Int J J Pharm Sci 86:991–996.
Pharm 341:1–19. Lo YL (2000) Phospholipids as multidrug resistance modulators of the transport of
Li P, Zhao LW, and Yalkowsky SH (1999d) Combined effect of cosolvent and cyclo- epirubicin in human intestinal epithelial Caco-2 cell layers and everted gut sacs of
dextrin on solubilization of nonpolar drugs. J Pharm Sci 88:1107–1111. rats. Biochem Pharmacol 60:1381–1390.
Li S, Pollock-Dove C, Dong LC, Chen J, Creasey AA, and Dai W-G (2012) Enhanced Lo YL (2003) Relationships between the hydrophilic-lipophilic balance values of
bioavailability of a poorly water-soluble weakly basic compound using a combina- pharmaceutical excipients and their multidrug resistance modulating effect in
tion approach of solubilization agents and precipitation inhibitors: a case study. Caco-2 cells and rat intestines. J Controlled Release 90:37–48.
Mol Pharm 9:1100–1108. Lockwood SF, O’Malley S, and Mosher GL (2003) Improved aqueous solubility of
Li SF, Doyle P, Metz S, Royce AE, and Serajuddin ATM (2005a) Effect of chloride ion crystalline astaxanthin (3,39-dihydroxy-beta, beta-carotene-4,49-dione) by Captisol
on dissolution of different salt forms of haloperidol, a model basic drug. J Pharm (sulfobutyl ether beta-cyclodextrin). J Pharm Sci 92:922–926.
Sci 94:2224–2231. Loe DW and Sharom FJ (1993) Interaction of multidrug-resistant Chinese hamster
Li SF, Wong SM, Sethia S, Almoazen H, Joshi YM, and Serajuddin ATM (2005b) ovary cells with amphiphiles. Br J Cancer 68:342–351.
Investigation of solubility and dissolution of a free base and two different salt forms Loftsson T and Brewster ME (1996) Pharmaceutical applications of cyclodextrins. 1.
as a function of pH. Pharm Res 22:628–635. Drug solubilization and stabilization. J Pharm Sci 85:1017–1025.
Li SM, Liu Y, Liu T, Zhao L, Zhao JH, and Feng NP (2011b) Development and in-vivo Loftsson T and Brewster ME (2010) Pharmaceutical applications of cyclodextrins:
assessment of the bioavailability of oridonin solid dispersions by the gas anti- basic science and product development. J Pharm Pharmacol 62:1607–1621.
solvent technique. Int J Pharm 411:172–177. Loftsson T and Brewster ME (2011) Pharmaceutical applications of cyclodextrins:
Li W, Yang YG, Tian YS, Xu XL, Chen Y, Mu LW, Zhang YQ, and Fang L (2011c) effects on drug permeation through biological membranes. J Pharm Pharmacol 63:
Preparation and in vitro/in vivo evaluation of revaprazan hydrochloride nano- 1119–1135.
suspension. Int J Pharm 408:157–162. Loftsson T and Duchêne D (2007) Cyclodextrins and their pharmaceutical applica-
Li Y and McClements DJ (2011) Inhibition of lipase-catalyzed hydrolysis of emulsi- tions. Int J Pharm 329:1–11.
fied triglyceride oils by low-molecular weight surfactants under simulated gas- Loftsson T, Frithriksdottir H, Thorisdottir S, and Stefansson E (1994) The effect of
trointestinal conditions. Eur J Pharm Biopharm 79:423–431. hydroxypropyl methylcellulose on the release of dexamethasone from aqueous 2-
Li Y, Paranawithana SR, Yoo JS, Ning SM, Ma BL, Lee MJ, Liu GT, and Yang CS hydroxypropyl-beta-cyclodextrin formulations. Int J Pharm 104:181–184.
(1992) Induction of liver microsomal cytochrome P-450 2B1 by dimethyl diphenyl Loftsson T, Hreinsdóttir D, and Másson M (2005) Evaluation of cyclodextrin solubi-
bicarboxylate in rats. Acta Pharmacol Sin 13:485–490. lization of drugs. Int J Pharm 302:18–28.
Lian T and Ho RJ (2001) Trends and developments in liposome drug delivery sys- Loftsson T, Hreinsdottir D, and Masson M (2007a) The complexation efficiency. J Incl
tems. J Pharm Sci 90:667–680. Phenom Macrocycl Chem 57:545–552.
Liddle J, Allen MJ, Borthwick AD, Brooks DP, Davies DE, Edwards RM, Exall AM, Loftsson T, Magnúsdóttir A, Másson M, and Sigurjónsdóttir JF (2002) Self-
Hamlett C, Irving WR, and Mason AM, et al. (2008) The discovery of GSK221149A: association and cyclodextrin solubilization of drugs. J Pharm Sci 91:2307–2316.
a potent and selective oxytocin antagonist. Bioorg Med Chem Lett 18:90–94. Loftsson T, Olafsdottir BJ, Frioriksdottir H, and Jonsdottir S (1993) Cyclodextrin
Liedholm H and Melander A (1986) Concomitant food intake can increase the bio- complexation of NSAIDS: physicochemical characteristics. Eur J Pharm Sci 1:
availability of propranolol by transient inhibition of its presystemic primary con- 95–101.
jugation. Clin Pharmacol Ther 40:29–36. Loftsson T, Vogensen SB, Brewster ME, and Konrádsdóttir F (2007b) Effects of
Lima AAN, Soares-Sobrinho JL, Silva JL, Corrêa-Júnior RAC, Lyra MAM, Santos cyclodextrins on drug delivery through biological membranes. J Pharm Sci 96:
FLA, Oliveira BG, Hernandes MZ, Rolim LA, and Rolim-Neto PJ (2011) The use of 2532–2546.
solid dispersion systems in hydrophilic carriers to increase benzonidazole solubil- Lohani S and Grant DJW (2006) Thermodynamics of polymorphism, in Poly-
ity. J Pharm Sci 100:2443–2451. morphism in the Pharmaceutical Industry (Hilfiker R ed) pp 21–42, WIiley-VCH,
Lin HS and Ho PC (2011) Preclinical pharmacokinetic evaluation of resveratrol tri- Weinheim, Germany.
methyl ether in sprague-dawley rats: the impacts of aqueous solubility, dose esca- Lohikangas L, Wilen M, Einarsson M, and Artursson P (1994) Effects of a new lipid-
lation, food and repeated dosing on oral bioavailability. J Pharm Sci 100:4491–4500. based drug-delivery system on the absorption of low-molecular-weight heparin
Lin HS, Leong WWY, Yang JA, Lee P, Chan SY, and Ho PC (2007a) Bio- (Fragmin) through monolayers of human intestinal epithelial Caco-2 Cells and
pharmaceutics of 13-cis-retinoic acid (isotretinoin) formulated with modified beta- after rectal administration to rabbits. Eur J Pharm Sci 1:297–305.
cyclodextrins. Int J Pharm 341:238–245. Lombardo D, Fauvel J, and Guy O (1980) Studies on the substrate specificity of
Lin JH (1999) Role of pharmacokinetics in the discovery and development of indi- a carboxyl ester hydrolase from human pancreatic juice. I. Action on carboxyl
navir. Adv Drug Deliv Rev 39:33–49. esters, glycerides and phospholipids. Biochim Biophys Acta 611:136–146.
Lin SL, Lachman L, Swartz CJ, and Huebner CF (1972) Preformulation in- Longwell A, Birss S, Keller N, and Moore D (1976) Effect of topically applied pilo-
vestigation. I. Relation of salt forms and biological activity of an experimental carpine on tear film pH. J Pharm Sci 65:1654–1657.
antihypertensive. J Pharm Sci 61:1418–1422. Lorenz W, Schmal A, Schult H, Lang S, Ohmann C, Weber D, Kapp B, Lüben L,
Lin Y, Shen Q, Katsumi H, Okada N, Fujita T, Jiang X, and Yamamoto A (2007b) and Doenicke A (1982) Histamine release and hypotensive reactions in dogs by
Effects of Labrasol and other pharmaceutical excipients on the intestinal transport solubilizing agents and fatty acids: analysis of various components in cremophor El
and absorption of rhodamine123, a P-glycoprotein substrate, in rats. Biol Pharm and development of a compound with reduced toxicity. Agents Actions 12:64–80.
Bull 30:1301–1307. Lowe ME (1997) Structure and function of pancreatic lipase and colipase. Annu Rev
Lindfors L, Forssén S, Skantze P, Skantze U, Zackrisson A, and Olsson U (2006) Nutr 17:141–158.
Amorphous drug nanosuspensions. 2. Experimental determination of bulk mono- Lu ATK, Frisella ME, and Johnson KC (1993) Dissolution modeling: factors affecting
mer concentrations. Langmuir 22:911–916. the dissolution rates of polydisperse powders. Pharm Res 10:1308–1314.
488 Williams et al.

Lu E, Rodriguez-Hornedo N, and Suryanarayanan R (2008) A rapid thermal method Masters K (1991) Spray Drying Handbook, Wiley, New York.
for cocrystal screening. CrystEngComm 10:665–668. Matsumoto T and Zografi G (1999) Physical properties of solid molecular dispersions
Lu GW, Hawley M, Smith M, Geiger BM, and Pfund W (2006) Characterization of of indomethacin with poly(vinylpyrrolidone) and poly(vinylpyrrolidone-co-vinyl-
a novel polymorphic form of celecoxib. J Pharm Sci 95:305–317. acetate) in relation to indomethacin crystallization. Pharm Res 16:1722–1728.
Lundin PDP, Bojrup M, Ljusberg-Wahren H, Weström BR, and Lundin S (1997) Matteucci ME, Brettmann BK, Rogers TL, Elder EJ, Williams RO 3rd, and Johnston
Enhancing effects of monohexanoin and two other medium-chain glyceride vehicles KP (2007) Design of potent amorphous drug nanoparticles for rapid generation of
on intestinal absorption of desmopressin (dDAVP). J Pharmacol Exp Ther 282: highly supersaturated media. Mol Pharm 4:782–793.
585–590. Mauludin R, Müller RH, and Keck CM (2009) Development of an oral rutin nano-
Lunkenheimer P, Schneider U, Brand R, and Loid A (2000) Glassy dynamics. Con- crystal formulation. Int J Pharm 370:202–209.
temp Phys 41:15–36. Maupas C, Moulari B, Béduneau A, Lamprecht A, and Pellequer Y (2011) Surfactant
Luo CF, Yuan M, Chen MS, Liu SM, Zhu L, Huang BY, Liu XW, and Xiong W (2011) dependent toxicity of lipid nanocapsules in HaCaT cells. Int J Pharm 411:136–141.
Pharmacokinetics, tissue distribution and relative bioavailability of puerarin solid Maurya D, Belgamwar V, and Tekade A (2010) Microwave induced solubility en-
lipid nanoparticles following oral administration. Int J Pharm 410:138–144. hancement of poorly water soluble atorvastatin calcium. J Pharm Pharmacol 62:
Luo Y, Chen DW, Ren LX, Zhao XL, and Qin J (2006) Solid lipid nanoparticles for 1599–1606.
enhancing vinpocetine’s oral bioavailability. J Control Release 114:53–59. McClements DJ (2012) Crystals and crystallization in oil-in-water emulsions:
MacGregor KJ, Embleton JK, Lacy JE, Perry EA, Solomon LJ, Seager H, and Pouton implications for emulsion-based delivery systems. Adv Colloid Interface Sci 174:
CW (1997) Influence of lipolysis on drug absorption from the gastro-intestinal 1–30.
tract. Adv Drug Deliv Rev 25:33–46. McConnell EL, Fadda HM, and Basit AW (2008) Gut instincts: explorations in in-
Macie CMG and Grant DJW (1986) Crystal-growth in pharmaceutical formulation. testinal physiology and drug delivery. Int J Pharm 364:213–226.
Pharm Int 7:233–237. McDevit WF and Long FA (1952) The activity coefficient of benzene in aqueous salt
Mackin L, Zanon R, Park JM, Foster K, Opalenik H, and Demonte M (2002) Quan- solutions. J Am Chem Soc 74:1773–1777.
tification of low levels (,10%) of amorphous content in micronized active batches McGinity JW, Maincent P, and Steinfink H (1984) Crystallinity and dissolution rate
using dynamic vapour sorption and isothermal microcalorimetry. Int J Pharm 231: of tolbutamide solid dispersions prepared by the melt method. J Pharm Sci 73:
227–236. 1441–1444.
Madieh S, Simone M, Wilson W, Mehra D, and Augsburger L (2007) Investigation of McIntosh MP, Leong N, Katneni K, Morizzi J, Shackleford DM, and Prankerd RJ
drug-porous adsorbent interactions in drug mixtures with selected porous (2010) Impact of chlorpromazine self-association on its apparent binding constants
adsorbents. J Pharm Sci 96:851–863. with cyclodextrins: effect of SBE(7)-beta-CD on the disposition of chlorpromazine in
Magnusdottir A, Masson M, and Loftsson T (2002) Self association and cyclodextrin the rat. J Pharm Sci 99:2999–3008.
solubilization of NSAIDs. J Incl Phenom Macrocycl Chem 44:213–218. McIntosh MP, Schwarting N, and Rajewski RA (2004) In vitro and in vivo evaluation
Magun AM, Brasitus TA, and Glickman RM (1985) Isolation of high density of a sulfobutyl ether beta-cyclodextrin enabled etomidate formulation. J Pharm Sci
lipoproteins from rat intestinal epithelial cells. J Clin Invest 75:209–218. 93:2585–2594.
Mahlin D, Ponnambalam S, Höckerfelt MH, and Bergström CAS (2011) Toward in McNamara DP and Amidon GL (1986) Dissolution of acidic and basic compounds
silico prediction of glass-forming ability from molecular structure alone: a screen- from the rotating disk: influence of convective diffusion and reaction. J Pharm Sci
ing tool in early drug development. Mol Pharm 8:498–506. 75:858–868.
Majerik V, Horváth G, Szokonya L, Charbit G, Badens E, Bosc N, and Teillaud E McNamara DP and Amidon GL (1988) Reaction plane approach for estimating the
(2007) Supercritical antisolvent versus coevaporation: preparation and character- effects of buffers on the dissolution rate of acidic drugs. J Pharm Sci 77:
ization of solid dispersions. Drug Dev Ind Pharm 33:975–983. 511–517.
Malcolmson C, Satra C, Kantaria S, Sidhu A, and Lawrence MJ (1998) Effect of oil on McNamara DP, Childs SL, Giordano J, Iarriccio A, Cassidy J, Shet MS, Mannion R,
the level of solubilization of testosterone propionate into nonionic oil-in-water O’Donnell E, and Park A (2006) Use of a glutaric acid cocrystal to improve oral
microemulsions. J Pharm Sci 87:109–116. bioavailability of a low solubility API. Pharm Res 23:1888–1897.
Malmsten M (2002) Surfactants and Polymers in Drug Delivery, Marcel Dekker, New Meanwell NA, Dennis RD, Roth HR, Rosenfeld MJ, Smith EC, Wright JJ, Buchanan
York. JO, Brassard CL, Gamberdella M, and Gillespie E, et al. (1992) Inhibitors of blood
Manca ML, Zaru M, Ennas G, Valenti D, Sinico C, Loy G, and Fadda AM (2005) platelet cAMP phosphodiesterase. 3. 1,3-dihydro-2H-imidazo[4,5-b]quinolin-2-one
Diclofenac-beta-cyclodextrin binary systems: physicochemical characterization derivatives with enhanced aqueous solubility. J Med Chem 35:2688–2696.
and in vitro dissolution and diffusion studies. AAPS PharmSciTech 6: Mehnert W and Mäder K (2001) Solid lipid nanoparticles: production, characteriza-
E464–E472. tion and applications. Adv Drug Deliv Rev 47:165–196.
Manjunath K and Venkateswarlu V (2005) Pharmacokinetics, tissue distribution and Mehuys E, Remon JP, Korst A, Van Bortel L, Mols R, Augustijns P, Porter CJH,
bioavailability of clozapine solid lipid nanoparticles after intravenous and intra- and Vervaet C (2005) Human bioavailability of propranolol from a matrix-in-
duodenal administration. J Control Release 107:215–228. cylinder system with a HPMC-Gelucire core. J Control Release 107:523–536.
Manjunath K and Venkateswarlu V (2006) Pharmacokinetics, tissue distribution and Mehuys E, Vervaet C, and Remon JP (2004) Hot-melt extruded ethylcellulose cyl-
bioavailability of nitrendipine solid lipid nanoparticles after intravenous and inders containing a HPMC-Gelucire core for sustained drug delivery. J Control
intraduodenal administration. J Drug Target 14:632–645. Release 94:273–280.
Manna L, Banchero M, Sola D, Ferri A, Ronchetti S, and Sicardi S (2007) Impreg- Meldrum FC and Sear RP (2008) Materials science: now you see them. Science 322:
nation of PVP microparticles with ketoprofen in the presence of supercritical CO2. 1802–1803.
J Supercrit Fluids 42:378–384. Mele A, Mendichi R, and Selva A (1998) Non-covalent associations of cyclo-
Mansbach CM 2nd and Gorelick F (2007) Development and physiological regulation maltooligosaccharides (cyclodextrins) with trans-beta-carotene in water: evidence
of intestinal lipid absorption. II. Dietary lipid absorption, complex lipid synthesis, for the formation of large aggregates by light scattering and NMR spectroscopy.
and the intracellular packaging and secretion of chylomicrons. Am J Physiol Carbohydr Res 310:261–267.
Gastrointest Liver Physiol 293:G645–G650. Mellaerts R, Aerts CA, Van Humbeeck J, Augustijns P, Van den Mooter G,
Mansbach CM II, Dowell RF, and Pritchett D (1991) Portal transport of absorbed and Martens JA (2007) Enhanced release of itraconazole from ordered mesoporous
lipids in rats. Am J Physiol 261:G530–G538. SBA-15 silica materials. Chem Commun (Camb) 13:1375–1377.
Mansbach CM and Siddiqi SA (2010) The biogenesis of chylomicrons. Annu Rev Mellaerts R, Mols R, Jammaer JAG, Aerts CA, Annaert P, Van Humbeeck J, Van den
Physiol 72:315–333. Mooter G, Augustijns P, and Martens JA (2008) Increasing the oral bioavailability
Mao C, Prasanth Chamarthy S, Byrn SR, and Pinal R (2006a) A calorimetric method of the poorly water soluble drug itraconazole with ordered mesoporous silica. Eur J
to estimate molecular mobility of amorphous solids at relatively low temperatures. Pharm Biopharm 69:223–230.
Pharm Res 23:2269–2276. Menegon RF, Blau L, Janzantti NS, Pizzolitto AC, Corrêa MA, Monteiro M,
Mao C, Chamarthy SP, and Pinal R (2006b) Time-dependence of molecular mobility and Chung MC (2011) A nonstaining and tasteless hydrophobic salt of chlorhex-
during structural relaxation and its impact on organic amorphous solids: an in- idine. J Pharm Sci 100:3130–3138.
vestigation based on a calorimetric approach. Pharm Res 23:1906–1917. Meng X, Mojaverian P, Doedée M, Lin E, Weinryb I, Chiang ST, and Kowey PR
Margolis J, McDonald J, Heuser R, Klinke P, Waksman R, Virmani R, Desai N, (2001) Bioavailability of amiodarone tablets administered with and without food in
and Hilton D (2007) Systemic nanoparticle paclitaxel (nab-paclitaxel) for in-stent healthy subjects. Am J Cardiol 87:432–435.
restenosis I (SNAPIST-I): a first-in-human safety and dose-finding study. Clin Menon R, Cefali E, Wilding I, Wray H, and Connor A (2009) The assessment of
Cardiol 30:165–170. human regional drug absorption of free acid and sodium salt forms of acipimox, in
Marsac PJ, Li T, and Taylor LS (2009) Estimation of drug-polymer miscibility and healthy volunteers, to direct modified release formulation strategy. Biopharm
solubility in amorphous solid dispersions using experimentally determined in- Drug Dispos 30:508–516.
teraction parameters. Pharm Res 26:139–151. Merisko-Liversidge E and Liversidge GG (2011) Nanosizing for oral and parenteral
Marsac PJ, Konno H, Rumondor ACF, and Taylor LS (2008) Recrystallization of drug delivery: a perspective on formulating poorly-water soluble compounds using
nifedipine and felodipine from amorphous molecular level solid dispersions con- wet media milling technology. Adv Drug Deliv Rev 63:427–440.
taining poly(vinylpyrrolidone) and sorbed water. Pharm Res 25:647–656. Merisko-Liversidge E, Liversidge GG, and Cooper ER (2003) Nanosizing: a formula-
Marsac PJ, Konno H, and Taylor LS (2006a) A comparison of the physical stability of tion approach for poorly-water-soluble compounds. Eur J Pharm Sci 18:113–120.
amorphous felodipine and nifedipine systems. Pharm Res 23:2306–2316. Merisko-Liversidge EM and Liversidge GG (2008) Drug nanoparticles: formulating
Marsac PJ, Shamblin SL, and Taylor LS (2006b) Theoretical and practical poorly water-soluble compounds. Toxicol Pathol 36:43–48.
approaches for prediction of drug-polymer miscibility and solubility. Pharm Res 23: Merisko-Liversidge E, Sarpotdar P, Bruno J, Hajj S, Wei L, Peltier N, Rake J, Shaw
2417–2426. JM, Pugh S, and Polin L, et al. (1996) Formulation and antitumor activity evalu-
Martin-Facklam M, Burhenne J, Ding R, Fricker R, Mikus G, Walter-Sack I, ation of nanocrystalline suspensions of poorly soluble anticancer drugs. Pharm Res
and Haefeli WE (2002a) Dose-dependent increase of saquinavir bioavailability by 13:272–278.
the pharmaceutic aid cremophor EL. Br J Clin Pharmacol 53:576–581. Merritt DA, Lynch MP, and King VL (2007) Pharmacokinetics of dirlotapide in the
Martin-Facklam M, Burhenne J, Ding R, Fricker R, Mikus G, Walter-Sack I, dog. J Vet Pharmacol Ther 30 (Suppl 1):24–32.
and Haefeli WE (2002b) Dose-dependent increase of saquinavir bioavailability by Messner M, Kurkov SV, Flavià-Piera R, Brewster ME, and Loftsson T (2011a) Self-
the pharmaceutic aid cremophor EL. Br J Clin Pharmacol 53:576–581. assembly of cyclodextrins: the effect of the guest molecule. Int J Pharm 408:235–247.
Strategies for Low Drug Solubility 489
Messner M, Kurkov SV, Jansook P, and Loftsson T (2010) Self-assembled cyclodex- Mooney KG, Mintun MA, Himmelstein KJ, and Stella VJ (1981) Dissolution kinetics
trin aggregates and nanoparticles. Int J Pharm 387:199–208. of carboxylic acids I. effect of pH under unbuffered conditions. J Pharm Sci 70:
Messner M, Kurkov SV, Maraver Palazón M, Álvarez Fernández B, Brewster ME, 13–22.
and Loftsson T (2011b) Self-assembly of cyclodextrin complexes: effect of tem- Moribe K, Tsutsumi S, Morishita S, Shinozaki H, Tozuka Y, Oguchi T,
perature, agitation and media composition on aggregation. Int J Pharm 419: and Yamamoto K (2005) Micronization of phenylbutazone by rapid expansion of
322–328. supercritical CO2 solution. Chem Pharm Bull (Tokyo) 53:1025–1028.
Millard JW, Alvarez-Núñez FA, and Yalkowsky SH (2002) Solubilization by cosol- Mörozowich W, Chulski T, Hamlin WE, Jones PM, Northam JI, Purmalis A,
vents: establishing useful constants for the log-linear model. Int J Pharm 245: and Wagner JG (1962) Relationship between in vitro dissolution rates, solubilities,
153–166. and LT50’s in mice of some salts of benzphetamine and etryptamine. J Pharm Sci
Miller DA, DiNunzio JC, Yang W, McGinity JW, and Williams RO 3rd (2008) En- 51:993–996.
hanced in vivo absorption of itraconazole via stabilization of supersaturation fol- Morris KR, Abramowitz R, Pinal R, Davis P, and Yalkowsky SH (1988) Solubility of
lowing acidic-to-neutral pH transition. Drug Dev Ind Pharm 34:890–902. aromatic pollutants in mixed-solvents. Chemosphere 17:285–298.
Miller DW, Batrakova EV, and Kabanov AV (1999) Inhibition of multidrug Morris KR, Fakes MG, Thakur AB, Newman AW, Singh AK, Venit JJ, Spagnuolo CJ,
resistance-associated protein (MRP) functional activity with pluronic block and Serajuddin ATM (1994) An integrated approach to the selection of optimal salt
copolymers. Pharm Res 16:396–401. form for a new drug candidate. Int J Pharm 105:209–217.
Millward MJ, Webster LK, Rischin D, Stokes KH, Toner GC, Bishop JF, Olver IN, Morris KR, Griesser UJ, Eckhardt CJ, and Stowell JG (2001) Theoretical approaches
Linahan BM, Linsenmeyer ME, and Woodcock DM (1998) Phase I trial of cremo- to physical transformations of active pharmaceutical ingredients during
phor EL with bolus doxorubicin. Clin Cancer Res 4:2321–2329. manufacturing processes. Adv Drug Deliv Rev 48:91–114.
Milton KA, Edwards G, Ward SA, Orme ML, and Breckenridge AM (1989) Phar- Morris KR, Knipp GT, and Serajuddin ATM (1992) Structural properties of poly-
macokinetics of halofantrine in man: effects of food and dose size. Br J Clin ethylene glycol-polysorbate 80 mixture, a solid dispersion vehicle. J Pharm Sci 81:
Pharmacol 28:71–77. 1185–1188.
Miro A, Quaglia F, Giannini L, Cappello B, and La Rotonda MI (2006) Drug/ Möschwitzer J, Achleitner G, Pomper H, and Müller RH (2004) Development of an
cyclodextrin solid systems in the design of hydrophilic matrices: a strategy to intravenously injectable chemically stable aqueous omeprazole formulation using
modulate drug delivery rate. Curr Drug Deliv 3:373–378. nanosuspension technology. Eur J Pharm Biopharm 58:615–619.
Mirtallo JM, Dasta JF, Kleinschmidt KC, and Varon J (2010) State of the art review: Mosharraf M and Nystrom C (1995) The effect of particle-size and shape on the
intravenous fat emulsions: current applications, safety profile, and clinical impli- surface specific dissolution rate of microsized practically insoluble drugs. Int J
cations. Ann Pharmacother 44:688–700. Pharm 122:35–47.
Mishra PR, Al Shaal L, Müller RH, and Keck CM (2009) Production and character- Mottu F, Laurent A, Rufenacht DA, and Doelker E (2000) Organic solvents for
ization of hesperetin nanosuspensions for dermal delivery. Int J Pharm 371: pharmaceutical parenterals and embolic liquids: a review of toxicity data. PDA J
182–189. Pharm Sci Technol 54:456–469.
Mitra AK and Mikkelson TJ (1982) Ophthalmic solution buffer systems. 1. The effect Mou D, Chen H, Wan J, Xu H, and Yang X (2011) Potent dried drug nanosuspensions
of buffer concentration on the ocular absorption of pilocarpine. Int J Pharm 10: for oral bioavailability enhancement of poorly soluble drugs with pH-dependent
219–229. solubility. Int J Pharm 413:237–244.
Miura H, Kanebako M, Shirai H, Nakao H, Inagi T, and Terada K (2010) Enhance- Mountfield RJ, Senepin S, Schleimer M, Walter I, and Bittner B (2000) Potential
ment of dissolution rate and oral absorption of a poorly water-soluble drug, K-832, inhibitory effects of formulation ingredients on intestinal cytochrome P450. Int J
by adsorption onto porous silica using supercritical carbon dioxide. Eur J Pharm Pharm 211:89–92.
Biopharm 76:215–221. Mu H and Høy CE (2004) The digestion of dietary triacylglycerols. Prog Lipid Res 43:
Miyake K, Arima H, Hirayama F, Yamamoto M, Horikawa T, Sumiyoshi H, Noda S, 105–133.
and Uekama K (2000) Improvement of solubility and oral bioavailability of rutin by Mudra DR and Borchardt RT (2010) Absorption barriers in the rat intestinal mucosa.
complexation with 2-hydroxypropyl-beta-cyclodextrin. Pharm Dev Technol 5: 3. Effects of polyethoxylated solubilizing agents on drug permeation and metabo-
399–407. lism. J Pharm Sci 99:1016–1027.
Miyake K, Arima H, Irie T, Hirayama F, and Uekama K (1999a) Enhanced absorp- Mueller EA, Kovarik JM, and Kutz K (1994a) Minor influence of a fat-rich meal on
tion of cyclosporin A by complexation with dimethyl-beta-cyclodextrin in bile duct- the pharmacokinetics of a new oral formulation of cyclosporine. Transplant Proc
cannulated and -noncannulated rats. Biol Pharm Bull 22:66–72. 26:2957–2958.
Miyake K, Irie T, Arima H, Hirayama F, Uekama K, Hirano M, and Okamaoto Y Mueller EA, Kovarik JM, van Bree JB, Grevel J, Lücker PW, and Kutz K (1994b)
(1999b) Characterization of itraconazole/2-hydroxypropyl-beta-cyclodextrin in- Influence of a fat-rich meal on the pharmacokinetics of a new oral formulation of
clusion complex in aqueous propylene glycol solution. Int J Pharm 179:237–245. cyclosporine in a crossover comparison with the market formulation. Pharm Res
Miyazaki S, Nakano M, and Arita T (1975) A comparison of solubility characteristics 11:151–155.
of free bases and hydrochloride salts of tetracycline antibiotics in hydrochloric acid Mueller EA, Kovarik JM, van Bree JB, Tetzloff W, Grevel J, and Kutz K (1994c)
solutions. Chem Pharm Bull (Tokyo) 23:1197–1204. Improved dose linearity of cyclosporine pharmacokinetics from a microemulsion
Miyazaki S, Oshiba M, and Nadai T (1980) Unusual solubility and dissolution be- formulation. Pharm Res 11:301–304.
havior of pharmaceutical hydrochloride salts in chloride-containing media. Int J Muhrer G, Meier U, Fusaro F, Albano S, and Mazzotti M (2006) Use of compressed
Pharm 6:77–85. gas precipitation to enhance the dissolution behavior of a poorly water-soluble
Miyazaki S, Oshiba M, and Nadai T (1981) Precaution on use of hydrochloride salts in drug: generation of drug microparticles and drug-polymer solid dispersions. Int J
pharmaceutical formulation. J Pharm Sci 70:594–596. Pharm 308:69–83.
Miyazaki T, Yoshioka S, and Aso Y (2006) Physical stability of amorphous acetanilide Mukerjee P (1979) Solubilization in aqueous micellar micelles, in Solution Chemistry
derivatives improved by polymer excipients. Chem Pharm Bull (Tokyo) 54: of Surfactants (Mittal KL ed) pp 153–174, Plenum Press, New York.
1207–1210. Muller RH, Becker R, Kruss B, and Peters K (1999), inventors Medac Gesellschafft
Miyazaki T, Yoshioka S, Aso Y, and Kawanishi T (2007) Crystallization rate of Fur Klinische Spezialpraparate, Hamburg, Germany, assignee. Pharmaceutical
amorphous nifedipine analogues unrelated to the glass transition temperature. Int nanosuspensions for medicament administration as systems with increases satu-
J Pharm 336:191–195. ration solubility and rate of solution. U.S. Patent, 5,858,410.
Miyazaki T, Yoshioka S, Aso Y, and Kojima S (2004) Ability of polyvinylpyrrolidone Müller RH and Bohm BHL (1998) Nanosuspensions, in Emulsions and Nano-
and polyacrylic acid to inhibit the crystallization of amorphous acetaminophen. suspensions for the Formulation of Poorly Soluble Drugs (Muller RH, Benita S,
J Pharm Sci 93:2710–2717. and Bohm B eds) pp 149–174, Medpharm, Stuttgart, Germany.
Moes JJ, Koolen SLW, Huitema ADR, Schellens JHM, Beijnen JH, and Nuijen B Müller RH, Gohla S, and Keck CM (2011a) State of the art of nanocrystals: special
(2011) Pharmaceutical development and preliminary clinical testing of an oral solid features, production, nanotoxicology aspects and intracellular delivery. Eur J
dispersion formulation of docetaxel (ModraDoc001). Int J Pharm 420:244–250. Pharm Biopharm 78:1–9.
Mohamed S, Tocher DA, and Price SL (2011) Computational prediction of salt and Müller RH, Jacobs C, and Kayser O (2001) Nanosuspensions as particulate drug
cocrystal structures: does a proton position matter? Int J Pharm 418:187–198. formulations in therapy. Rationale for development and what we can expect for the
Mohsin K, Long MA, and Pouton CW (2009) Design of lipid-based formulations for future. Adv Drug Deliv Rev 47:3–19.
oral administration of poorly water-soluble drugs: precipitation of drug after dis- Müller RH and Junghanns JAW (1998) Drug nanocrystals/nanosuspensions for the
persion of formulations in aqueous solution. J Pharm Sci 98:3582–3595. delivery of poorly soluble drugs, in Nanoparticulates as drug carriers (Torchilin VP
Moneghini M, De Zordi N, Solinas D, Macchiavelli S, and Princivalle F (2010) ed) pp 307–328, Imperial College Press London.
Characterization of solid dispersions of itraconazole and vitamin E TPGS prepared Müller RH, Mäder K, and Gohla S (2000) Solid lipid nanoparticles (SLN) for con-
by microwave technology. Future Med Chem 2:237–246. trolled drug delivery: a review of the state of the art. Eur J Pharm Biopharm 50:
Moneghini M, Kikic I, Voinovich D, Perissutti B, and Filipovic-Grcic J (2001) Pro- 161–177.
cessing of carbamazepine-PEG 4000 solid dispersions with supercritical carbon Müller RH, Radtke M, and Wissing SA (2002) Nanostructured lipid matrices for
dioxide: preparation, characterisation, and in vitro dissolution. Int J Pharm 222: improved microencapsulation of drugs. Int J Pharm 242:121–128.
129–138. Müller RH, Runge S, Ravelli V, Mehnert W, Thünemann AF, and Souto EB (2006)
Moneghini M, Zingone G, and De Zordi N (2009) Influence of the microwave tech- Oral bioavailability of cyclosporine: solid lipid nanoparticles (SLN) versus drug
nology on the physical-chemical properties of solid dispersion with nimesulide. nanocrystals. Int J Pharm 317:82–89.
Powder Technol 195:259–263. Müller RH, Shegokar R, and Keck CM (2011b) 20 years of lipid nanoparticles (SLN
Morissette SL, Almarsson O, Peterson ML, Remenar JF, Read MJ, Lemmo AV, Ellis and NLC): present state of development and industrial applications. Curr Drug
S, Cima MJ, and Gardner CR (2004) High-throughput crystallization: polymorphs, Discov Technol 8:207–227.
salts, co-crystals and solvates of pharmaceutical solids. Adv Drug Deliv Rev 56: Müllertz A, Fatouros DG, Smith JR, Vertzoni M, and Reppas C (2012) Insights into
275–300. intermediate phases of human intestinal fluids visualized by atomic force mi-
Monkhouse DC and Lach JL (1972a) Use of adsorbents in enhancement of drug croscopy and cryo-transmission electron microscopy ex vivo. Mol Pharm 9:237–247.
dissolution. I. J Pharm Sci 61:1430–1435. Müllertz A, Ogbonna A, Ren S, and Rades T (2010) New perspectives on lipid and
Monkhouse DC and Lach JL (1972b) Use of adsorbents in enhancement of drug surfactant based drug delivery systems for oral delivery of poorly soluble drugs.
dissolution. II. J Pharm Sci 61:1435–1441. J Pharm Pharmacol 62:1622–1636.
490 Williams et al.

Mura P, Adragna E, Rabasco AM, Moyano JR, Pérez-Martìnez JI, Arias MJ, ibuprofen binary solid dispersions with poloxamer 188. Int J Pharm 343:
and Ginés JM (1999) Effects of the host cavity size and the preparation method on 228–237.
the physicochemical properties of ibuproxam-cyclodextrin systems. Drug Dev Ind Newman A, Knipp G, and Zografi G (2012) Assessing the performance of amorphous
Pharm 25:279–287. solid dispersions. J Pharm Sci 101:1355–1377.
Mura P, Bettinetti GP, Cirri M, Maestrelli F, Sorrenti M, and Catenacci L (2005) Ngai KL (1998) Correlation between the secondary beta-relaxation time at Tg with
Solid-state characterization and dissolution properties of naproxen-arginine- the Kohlrausch exponent of the primary alpha relaxation or the fragility of glass-
hydroxypropyl-beta-cyclodextrin ternary system. Eur J Pharm Biopharm 59:99–106. forming materials. Phys Rev E 57:7346–7349.
Mura P, Bettinetti GP, Manderioli A, Faucci MT, Bramanti G, and Sorrenti M (1998) Ngai KL, Capaccioli S, Shinyashiki N, and Thayyil MS (2010) Recent progress in
Interactions of ketoprofen and ibuprofen with beta-cyclodextrins in solution and in understanding relaxation in complex systems. J Non-Cryst Solids 356:535–541.
the solid state. Int J Pharm 166:189–203. Ngai KL, Lunkenheimer P, Leon C, Schneider U, Brand R, and Loidl A (2001) Nature
Mura P, Faucci MT, and Bettinetti GP (2001a) The influence of polyvinylpyrrolidone and properties of the Johari-Goldstein beta-relaxation in the equilibrium liquid
on naproxen complexation with hydroxypropyl-beta-cyclodextrin. Eur J Pharm Sci state of a class of glass-formers. J Chem Phys 115:1405–1413.
13:187–194. Nguyen TH, Hanley T, Porter CJH, Larson I, and Boyd BJ (2010) Phytantriol and
Mura P, Faucci MT, Manderioli A, and Bramanti G (2001b) Multicomponent systems glyceryl monooleate cubic liquid crystalline phases as sustained-release oral drug
of econazole with hydroxyacids and cyclodextrins. J Incl Phenom Macrocycl Chem delivery systems for poorly water-soluble drugs. II. In-vivo evaluation. J Pharm
39:131–138. Pharmacol 62:856–865.
Mura P, Maestrelli F, and Cirri M (2003) Ternary systems of naproxen with Ni N, Sanghvi T, and Yalkowsky SH (2002) Solubilization and preformulation of
hydroxypropyl-beta-cyclodextrin and aminoacids. Int J Pharm 260:293–302. carbendazim. Int J Pharm 244:99–104.
Murdam S and Florence AT (2000) Non-aqueous solutions and suspensions as Nicklasson M and Brodin A (1984) The relationship between intrinsic dissolution
sustained-release injectable formulations, in Sustained-Release Injectable Products rates and solubilities in the water ethanol binary solvent system. Int J Pharm 18:
(Senior J and Radomsky M eds) pp 71–107, Interpharm Press, Denver, CO. 149–156.
Murdande SB, Pikal MJ, Shanker RM, and Bogner RH (2010a) Solubility advantage Nicklasson M, Brodin A, and Nyqvist H (1981) Studies on the relationship between
of amorphous pharmaceuticals: I. A thermodynamic analysis. J Pharm Sci 99: solubility and intrinsic rate of dissolution as a function of pH. Acta Pharm Suec 18:
1254–1264. 119–128.
Murdande SB, Pikal MJ, Shanker RM, and Bogner RH (2010b) Solubility advantage Nicolaides E, Symillides M, Dressman JB, and Reppas C (2001) Biorelevant disso-
of amorphous pharmaceuticals: II. Application of quantitative thermodynamic lution testing to predict the plasma profile of lipophilic drugs after oral adminis-
relationships for prediction of solubility enhancement in structurally diverse in- tration. Pharm Res 18:380–388.
soluble pharmaceuticals. Pharm Res 27:2704–2714. Nielloud F and Marti-Mestres G (2000) Pharmaceutical Emulsions and Suspensions,
Murdande SB, Pikal MJ, Shanker RM, and Bogner RH (2011a) Aqueous solubility of Marcel Dekker Inc, Basel, Switzerland.
crystalline and amorphous drugs: challenges in measurement. Pharm Dev Technol Nimmerfall F and Rosenthaler J (1980) Dependence of area under the curve on
16:187–200. proquazone particle size and in vitro dissolution rate. J Pharm Sci 69:605–607.
Murdande SB, Pikal MJ, Shanker RM, and Bogner RH (2011b) Solubility advantage Niot I and Besnard P (2004) Intestinal fat absorption: roles of intracellular lipid-
of amorphous pharmaceuticals. III. Is maximum solubility advantage experimen- binding proteins and peroxisome proliferator-activated receptors, in Cellular Pro-
tally attainable and sustainable? J Pharm Sci 100:4349–4356. teins and Their Fatty Acids in Health and Disease pp 359–381, Wiley-VCH,
Myers RA and Stella VJ (1992) Systemic bioavailability of penclomedine (NSC- Weinheim, Germany.
338720) from oil-in-water emulsions administered intraduodenally to rats. Int J Niot I, Poirier H, Tran TT, and Besnard P (2009) Intestinal absorption of long-chain
Pharm 78:217–226. fatty acids: evidence and uncertainties. Prog Lipid Res 48:101–115.
Myers SL and Shively ML (1993) Solid-state emulsions: the effects of maltodextrin on Nishino M, Enachescu C, Miyashita S, Rikvold PA, Boukheddaden K, and Varret F
microcrystalline aging. Pharm Res 10:1389–1391. (2011) Macroscopic nucleation phenomena in continuum media with long-range
Na GC, Yuan BO, Stevens HJ Jr, Weekley BS, and Rajagopalan N (1999) Cloud point interactions. Scientific Reports 1 DOI: 10.1038/srep00162.
of nonionic surfactants: modulation with pharmaceutical excipients. Pharm Res 16: Niwa T, Miura S, and Danjo K (2011a) Design of dry nanosuspension with highly
562–568. spontaneous dispersible characteristics to develop solubilized formulation for
Nagapudi K and Jona J (2008) Amorphous active pharmaceutical ingredients in poorly water-soluble drugs. Pharm Res 28:2339–2349.
preclinical studies. Current Bioactive Compounds 4:213–224. Niwa T, Miura S, and Danjo K (2011b) Universal wet-milling technique to prepare
Nagarsenker MS, Meshram RN, and Ramprakash G (2000) Solid dispersion of oral nanosuspension focused on discovery and preclinical animal studies: de-
hydroxypropyl beta-cyclodextrin and ketorolac: enhancement of in-vitro dissolution velopment of particle design method. Int J Pharm 405:218–227.
rates, improvement in anti-inflammatory activity and reduction in ulcerogenicity Nokhodchi A, Bolourtchian N, and Dinarvand R (2003) Crystal modification of phe-
in rats. J Pharm Pharmacol 52:949–956. nytoin using different solvents and crystallization conditions. Int J Pharm 250:
Nakach M, Authelin JR, Chamayou A, and Dodds J (2004) Comparison of various 85–97.
milling technologies for grinding pharmaceutical powders. Int J Miner Process 74: Nordskog BK, Phan CT, Nutting DF, and Tso P (2001) An examination of the factors
S173–S181. affecting intestinal lymphatic transport of dietary lipids. Adv Drug Deliv Rev 50:
Nakai Y, el-Said Aboutaleb A, Yamamoto K, Saleh SI, and Ahmed MO (1990) Study 21–44.
of the interaction of clobazam with cyclodextrins in solution and in the solid state. Noyes AA and Whitney WR (1897) The rate of solution of solid substances in their
Chem Pharm Bull (Tokyo) 38:728–732. own solutions. J Am Chem Soc 19:930–934.
Nakanishi K, Masada M, Nadai T, and Miyajima K (1989) Effect of the interaction of Nykamp G, Carstensen U, and Müller BW (2002) Jet milling: a new technique for
drug-beta-cyclodextrin complex with bile salts on the drug absorption from rat microparticle preparation. Int J Pharm 242:79–86.
small intestinal lumen. Chem Pharm Bull (Tokyo) 37:211–214. Nystrom C (1998) Dissolution properties of poorly soluble drugs: Theoretical back-
Nambu N, Shimoda M, Takahashi Y, Ueda H, and Nagai T (1978) Bioavailability of ground and possibilities to improve dissolution behaviour, in Emulsions and
powdered inclusion compounds of nonsteroidal anti-inflammatory drugs with b2 Nanosuspensions for the Formulation of Poorly Soluble Drugs (Muller RH, Benita
cyclodextrin in rabbits and dogs. Chem Pharm Bull (Tokyo) 26:2952–2956. S, and Bohm B eds) pp 143–148, Medpharm, Stuttgart, Germany.
Nandi I, Bateson M, Bari M, and Joshi HN (2003) Synergistic effect of PEG-400 and O’Connor KM and Corrigan OI (2001) Preparation and characterisation of a range of
cyclodextrin to enhance solubility of progesterone. AAPS PharmSciTech 4:E1. diclofenac salts. Int J Pharm 226:163–179.
Narazaki R, Sanghvi R, and Yalkowsky SH (2007a) Estimation of drug precipitation O’Driscoll CM (2002) Lipid-based formulations for intestinal lymphatic delivery. Eur
upon dilution of pH-controlled formulations. Mol Pharm 4:550–555. J Pharm Sci 15:405–415.
Narazaki R, Sanghvi R, and Yalkowsky SH (2007b) Estimation of drug precipitation Offerdahl TJ, Salsbury JS, Dong ZD, Grant DJW, Schroeder SA, Prakash I, Gorman
upon dilution of pH-cosolvent solubilized formulations. Chem Pharm Bull (Tokyo) EM, Barich DH, and Munson EJ (2005) Quantitation of crystalline and amorphous
55:1203–1206. forms of anhydrous neotame using 13C CPMAS NMR spectroscopy. J Pharm Sci
Nassar MN, Nesarikar VN, Lozano R, Parker WL, Huang YD, Palaniswamy V, Xu 94:2591–2605.
WW, and Khaselev N (2004) Influence of formaldehyde impurity in polysorbate 80 Oh DH, Kang JH, Kim DW, Lee B-J, Kim JO, Yong CS, and Choi H-G (2011) Com-
and PEG-300 on the stability of a parenteral formulation of BMS-204352: identi- parison of solid self-microemulsifying drug delivery system (solid SMEDDS) pre-
fication and control of the degradation product. Pharm Dev Technol 9:189–195. pared with hydrophilic and hydrophobic solid carrier. Int J Pharm 420:412–418.
Neervannan S (2006) Preclinical formulations for discovery and toxicology: physico- Oh KT, Bronich TK, and Kabanov AV (2004) Micellar formulations for drug delivery
chemical challenges. Expert Opin Drug Metab Toxicol 2:715–731. based on mixtures of hydrophobic and hydrophilic Pluronic block copolymers. J
Nehm SJ, Rodriguez-Spong B, and Rodriguez-Hornedo N (2006) Phase solubility Control Release 94:411–422.
diagrams of cocrystals are explained by solubility product and solution complexa- Ohara T, Kitamura S, Kitagawa T, and Terada K (2005) Dissolution mechanism of
tion. Cryst Growth Des 6:592–600. poorly water-soluble drug from extended release solid dispersion system with
Nekkanti V, Pillai R, Venkateshwarlu V, and Harisudhan T (2009) Development and ethylcellulose and hydroxypropylmethylcellulose. Int J Pharm 302:95–102.
characterization of solid oral dosage form incorporating candesartan nanoparticles. Ohgaki K, Hirokawa N, and Ueda M (1992) Heterogeneity in aqueous-solutions:
Pharm Dev Technol 14:290–298. electron-microscopy of citric-acid solutions. Chem Eng Sci 47:1819–1823.
Nelson E (1957) Solution rate of theophylline salts and effects from oral adminis- Okimoto K, Miyake M, Ibuki R, Yasumura M, Ohnishi N, and Nakai T (1997) Dis-
tration. J Pharm Sci 46:607–614. solution mechanism and rate of solid dispersion particles of nilvadipine with
Nelson E, Knoechel EL, Hamlin WE, and Wagner JG (1962) Influence of the ab- hydroxypropylmethylcellulose. Int J Pharm 159:85–93.
sorption rate of tolbutamide on the rate of decline of blood sugar levels in normal Okimoto K, Rajewski RA, and Stella VJ (1999) Release of testosterone from an os-
humans. J Pharm Sci 51:509–514. motic pump tablet utilizing (SBE)7m-beta-cyclodextrin as both a solubilizing and
Nema S and Brendel RJ (2011) Excipients and their role in approved injectable prod- an osmotic pump agent. J Control Release 58:29–38.
ucts: current usage and future directions. PDA J Pharm Sci Technol 65:287–332. Okimoto K, Rajewski RA, Uekama K, Jona JA, and Stella VJ (1996) The interaction
Nema S, Washkuhn RJ, and Brendel RJ (1997) Excipients and their use in injectable of charged and uncharged drugs with neutral (HP-beta-CD) and anionically
products. PDA J Pharm Sci Technol 51:166–171. charged (SBE7-beta-CD) beta-cyclodextrins. Pharm Res 13:256–264.
Newa M, Bhandari KH, Li DX, Kwon TH, Kim JA, Yoo BK, Woo JS, Lyoo WS, Yong Okonogi S, Oguchi T, Yonemochi E, Puttipipatkhachorn S, and Yamamoto K (1997)
CS, and Choi HG (2007) Preparation, characterization and in vivo evaluation of Improved dissolution of ofloxacin via solid dispersion. Int J Pharm 156:175–180.
Strategies for Low Drug Solubility 491
Oliveira HC, Nilausen K, Meinertz H, and Quintão EC (1993) Cholesteryl esters in Passerini N, Albertini B, Perissutti B, and Rodriguez L (2006) Evaluation of melt
lymph chylomicrons: contribution from high density lipoprotein transferred from granulation and ultrasonic spray congealing as techniques to enhance the disso-
plasma into intestinal lymph. J Lipid Res 34:1729–1736. lution of praziquantel. Int J Pharm 318:92–102.
Omar L, El-Barghouthi MI, Masoud NA, Abdoh AA, Al Omari MM, Zughul MB, Patel AR and Vavia PR (2006) Effect of hydrophilic polymer on solubilization of
and Badwan AA (2007) Inclusion complexation of loratadine with natural and fenofibrate by cyclodextrin complexation. J Incl Phenom Macrocycl Chem 56:
modified cyclodextrins: phase solubility and thermodynamic studies. J Solution 247–251.
Chem 36:605–616. Patel JP and Brocks DR (2009) The effect of oral lipids and circulating lipoproteins on
Ong JTH and Manoukian E (1988) Micellar solubilization of timobesone acetate in the metabolism of drugs. Expert Opin Drug Metab Toxicol 5:1385–1398.
aqueous and aqueous propylene glycol solutions of nonionic surfactants. Pharm Patel JP, Korashy HM, El-Kadi AO, and Brocks DR (2010) Effect of bile and lipids on
Res 5:704–708. the stereoselective metabolism of halofantrine by rat everted-intestinal sacs.
Onizuka M, Flatebø T, and Nicolaysen G (1997) Lymph flow pattern in the intact Chirality 22:275–283.
thoracic duct in sheep. J Physiol 503:223–234. Patil P and Paradkar A (2006) Porous polystyrene beads as carriers for self-
Orlowski S, Selosse MA, Boudon C, Micoud C, Mir LM, Belehradek J Jr, and Garrigos emulsifying system containing loratadine. AAPS PharmSciTech 7:E28.
M (1998) Effects of detergents on P-glycoprotein atpase activity: differences in Patravale VB, Date AA, and Kulkarni RM (2004) Nanosuspensions: a promising drug
perturbations of basal and verapamil-dependent activities. Cancer Biochem Bio- delivery strategy. J Pharm Pharmacol 56:827–840.
phys 16:85–110. Patt WC, Cheng XM, Repine JT, Lee C, Reisdorph BR, Massa MA, Doherty AM,
Orsenigo MN, Tosco M, Zoppi S, and Faelli A (1990) Characterization of basolateral Welch KM, Bryant JW, and Flynn MA, et al. (1999) Butenolide endothelin
membrane Na/H antiport in rat jejunum. Biochim Biophys Acta 1026:64–68. antagonists with improved aqueous solubility. J Med Chem 42:2162–2168.
Otaegui D, Rodríguez-Gascón A, Zubia A, Cossío FP, and Pedraz JL (2009) Phar- Patterson JE, James MB, Forster AH, and Rades T (2008) Melt extrusion and spray
macokinetics and tissue distribution of Kendine 91, a novel histone deacetylase drying of carbamazepine and dipyridamole with polyvinylpyrrolidone/vinyl acetate
inhibitor, in mice. Cancer Chemother Pharmacol 64:153–159. copolymers. Drug Dev Ind Pharm 34:95–106.
Oteroespinar FJ, Anguianoigea S, Garciagonzalez N, Vilajato JL, and Blancomendez Patton JS and Carey MC (1979) Watching fat digestion. Science 204:145–148.
J (1991) Oral bioavailability of naproxen-beta-cyclodextrin inclusion compound. Int Patton JS, Vetter RD, Hamosh B, Borgstroem B, Lindstrom MC, and Carey MC
J Pharm 75:37–44. (1984) The light microscopy of triglyceride digestion. Food Microstruct 4:29–41.
Otsuka M, Onoe M, and Matsuda Y (1993) Hygroscopic stability and dissolution Paudel A, Van Humbeeck J, and Van den Mooter G (2010) Theoretical and experi-
properties of spray-dried solid dispersions of furosemide with Eudragit. J Pharm mental investigation on the solid solubility and miscibility of naproxen in
Sci 82:32–38. poly(vinylpyrrolidone). Mol Pharm 7:1133–1148.
Ozaki S, Minamisono T, Yamashita T, Kato T, and Kushida I (2012) Supersaturation- Paulekuhn GS, Dressman JB, and Saal C (2007) Trends in active pharmaceutical
nucleation behavior of poorly soluble drugs and its impact on the oral absorption of ingredient salt selection based on analysis of the Orange Book database. J Med
drugs in thermodynamically high-energy forms. J Pharm Sci 101:214–222. Chem 50:6665–6672.
Padden B, Miller JM, Robbins T, Zocharski PD, Prasad L, Spence JK, Pazdur R, Kudelka AP, Kavanagh JJ, Cohen PR, and Raber MN (1993) The taxoids:
and LaFountaine J (2011) Amorphous solid dispersions as enabling formulations paclitaxel (Taxol) and docetaxel (Taxotere). Cancer Treat Rev 19:351–386.
for discovery and early development. Am Pharm Rev 1:66–73. Pedersen M (1997) The bioavailability difference between genuine cyclodextrin in-
Padrela L, Rodrigues MA, Velaga SP, Matos HA, and de Azevedo EG (2009) For- clusion complexes and freeze-dried or ground drug cyclodextrin samples may be
mation of indomethacin-saccharin cocrystals using supercritical fluid technology. due to supersaturation differences. Drug Dev Ind Pharm 23:331–335.
Eur J Pharm Sci 38:9–17. Pedersen M, Edelsten M, Nielsen VF, Scarpellini A, Skytte S, and Slot C (1993)
Pajula K, Taskinen M, Lehto VP, Ketolainen J, and Korhonen O (2010) Predicting Formation and antimycotic effect of cyclodextrin inclusion complexes of econazole
the formation and stability of amorphous small molecule binary mixtures from and miconazole. Int J Pharm 90:247–254.
computationally determined Flory-Huggins interaction parameter and phase dia- Peeters J, Neeskens P, Tollenaere JP, Van Remoortere P, and Brewster ME (2002)
gram. Mol Pharm 7:795–804. Characterization of the interaction of 2-hydroxypropyl-beta-cyclodextrin with
Palakodaty S and York P (1999) Phase behavioral effects on particle formation pro- itraconazole at pH 2, 4, and 7. J Pharm Sci 91:1414–1422.
cesses using supercritical fluids. Pharm Res 16:976–985. Peltonen L and Hirvonen J (2010) Pharmaceutical nanocrystals by nanomilling:
Palay SL and Karlin LJ (1959) An electron microscopic study of the intestinal villus. critical process parameters, particle fracturing and stabilization methods. J Pharm
II. The pathway of fat absorption. J Biophys Biochem Cytol 5:373–384. Pharmacol 62:1569–1579.
Palin KJ and Wilson CG (1984) The effect of different oils on the absorption of Perng CY, Kearney AS, Palepu NR, Smith BR, and Azzarano LM (2003) Assessment
probucol in the rat. J Pharm Pharmacol 36:641–643. of oral bioavailability enhancing approaches for SB-247083 using flow-through cell
Palucki M, Higgins JD, Kwong E, and Templeton AC (2010) Strategies at the in- dissolution testing as one of the screens. Int J Pharm 250:147–156.
terface of drug discovery and development: early optimization of the solid state Perry CS, Charman SA, Prankerd RJ, Chiu FCK, Scanlon MJ, Chalmers D,
phase and preclinical toxicology formulation for potential drug candidates. J Med and Charman WN (2006) The binding interaction of synthetic ozonide anti-
Chem 53:5897–5905. malarials with natural and modified beta-cyclodextrins. J Pharm Sci 95:146–158.
Pan SH, Lopez RR Jr, Sher LS, Hoffman AL, Podesta LG, Makowka L, and Rosenthal Peters K, Leitzke S, Diederichs JE, Borner K, Hahn H, Müller RH, and Ehlers S
P (1996) Enhanced oral cyclosporine absorption with water-soluble vitamin E early (2000) Preparation of a clofazimine nanosuspension for intravenous use and
after liver transplantation. Pharmacotherapy 16:59–65. evaluation of its therapeutic efficacy in murine Mycobacterium avium infection.
Pan XH, Julian T, and Augsburger L (2008) Increasing the dissolution rate of a low- J Antimicrob Chemother 45:77–83.
solubility drug through a crystalline-amorphous transition: a case study with Peterson ML, Collier EA, Hickey MB, Guzman H, and Almarsson O (2010) Multi-
indomethicin. Drug Dev Ind Pharm 34:221–231. component pharmaceutical crystalline phases: engineering for performance, in
Pandit JK, Gupta SK, Gode KD, and Mishra B (1984) Effect of crystal form on the Organic Crystal Engineering: Frontiers in Crystal Engineering (Tiekink ERT,
oral absorption of phenylbutazone. Int J Pharm 21:129–132. Vittal J, and Zaworotko MJ eds) Wiley, New York.
Papadimitriou SA, Bikiaris D, and Avgoustakis K (2008) Microwave-induced en- Phan CT and Tso P (2001) Intestinal lipid absorption and transport. Front Biosci 6:
hancement of the dissolution rate of poorly water-soluble tibolone from poly(eth- D299–D319.
ylene glycol) solid dispersions. J Appl Polym Sci 108:1249–1258. Lee YHP, Sathigari S, Jean Lin YJ, Ravis WR, Chadha G, Parsons DL, Rangari VK,
Papp M, Rohlich P, Rusznyak I, and Toro I (1962) An electron microscopic study of Wright N, and Babu RJ (2009) Gefitinib-cyclodextrin inclusion complexes: physico-
the central lacteal in the intestinal villus of the cat. Z Zellforsch Mikrosk Anat 57: chemical characterization and dissolution studies. Drug Dev Ind Pharm 35:
475–486. 1113–1120.
Pardeike J, Hommoss A, and Müller RH (2009) Lipid nanoparticles (SLN, NLC) in Phillips EM and Stella VJ (1993) Rapid expansion from supercritical solutions: ap-
cosmetic and pharmaceutical dermal products. Int J Pharm 366:170–184. plication to pharmaceutical processes. Int J Pharm 94:1–10.
Pardeike J and Müller RH (2010) Nanosuspensions: a promising formulation for the Piel G, Evrard B, Fillet M, Llabres G, and Delattre L (1998) Development of a non-
new phospholipase A2 inhibitor PX-18. Int J Pharm 391:322–329. surfactant parenteral formulation of miconazole by the use of cyclodextrins. Int J
Park JH, Chi SC, Lee WS, Lee WM, Koo YB, Yong CS, Choi HG, and Woo JS (2009) Pharm 169:15–22.
Toxicity studies of cremophor-free paclitaxel solid dispersion formulated by a su- Piel G, Evrard B, Van Hees T, and Delattre L (1999) Comparison of the IV phar-
percritical antisolvent process. Arch Pharm Res 32:139–148. macokinetics in sheep of miconazole-cyclodextrin solutions and a micellar solution.
Park JW and Song HJ (1989) Association of anionic surfactants with beta- Int J Pharm 180:41–45.
cyclodextrin: fluorescence-probed studies on the 1-1 and 1-2 complexation. J Piel G, Pirotte B, Delneuville I, Neven P, Llabres G, Delarge J, and Delattre L (1997)
Phys Chem 93:6454–6458. Study of the influence of both cyclodextrins and L-lysine on the aqueous solubility
Parshad H, Frydenvang K, Liljefors T, and Larsen CS (2002) Correlation of aqueous of nimesulide; isolation and characterization of nimesulide-L-lysine-cyclodextrin
solubility of salts of benzylamine with experimentally and theoretically derived complexes. J Pharm Sci 86:475–480.
parameters: a multivariate data analysis approach. Int J Pharm 237:193–207. Plakkot S, de Matas M, York P, Saunders M, and Sulaiman B (2011) Comminution of
Parshad H, Frydenvang K, Liljefors T, Sorensen HO, and Larsen C (2004) Aqueous ibuprofen to produce nano-particles for rapid dissolution. Int J Pharm 415:
solubility study of salts of benzylamine derivatives and p-substituted benzoic acid 307–314.
derivatives using X-ray crystallographic analysis. Int J Pharm 269:157–168. Pohl J, Ring A, Ehehalt R, Schulze-Bergkamen H, Schad A, Verkade P,
Partyka M, Au BH, and Evans CH (2001) Cyclodextrins as phototoxicity inhibitors in and Stremmel W (2004) Long-chain fatty acid uptake into adipocytes depends on
drug formulations: studies on model systems involving naproxen and beta- lipid raft function. Biochemistry 43:4179–4187.
cyclodextrin. J Photochem Photobiol a. Chemistry 140:67–74. Pohl J, Ring A, and Stremmel W (2002) Uptake of long-chain fatty acids in
Paruta AN, Sciarrone BJ, and Lordi NG (1964) Solubility of salicylic acid as a func- HepG2 cells involves caveolae: analysis of a novel pathway. J Lipid Res 43:
tion of dielectric constant. J Pharm Sci 53:1349–1353. 1390–1399.
Pasini CE and Indelicato JM (1992) Pharmaceutical properties of loracarbef: the Pole DL (2008) Physical and biological considerations for the use of nonaqueous
remarkable solution stability of an oral 1-carba-1-dethiacephalosporin antibiotic. solvents in oral bioavailability enhancement. J Pharm Sci 97:1071–1088.
Pharm Res 9:250–254. Popovici RF, Seftel EM, Mihai GD, Popovici E, and Voicu VA (2011) Controlled drug
Pasquali I, Bettini R, and Giordano F (2006) Solid-state chemistry and particle en- delivery system based on ordered mesoporous silica matrices of captopril as
gineering with supercritical fluids in pharmaceutics. Eur J Pharm Sci 27:299–310. angiotensin-converting enzyme inhibitor drug. J Pharm Sci 100:704–714.
492 Williams et al.

Porter CJH, Charman SA, Williams RD, Bakalova MV, and Charman WN (1996) Quan YS, Hattori K, Lundborg E, Fujita T, Murakami M, Muranishi S,
Evaluation of emulsifiable glasses for the oral administration of cyclosporin in and Yamamoto A (1998) Effectiveness and toxicity screening of various absorption
beagle dogs. Int J Pharm 141:227–237. enhancers using Caco-2 cell monolayers. Biol Pharm Bull 21:615–620.
Porter CJH, Anby MU, Warren DB, Williams HD, Benameur H, and Pouton CW Quintão EC, Drewiacki A, Stechhaln K, de Faria EC, and Sipahi AM (1979) Origin of
(2011) Lipid based formulations: exploring the link between in vitro supersatura- cholesterol transported in intestinal lymph: studies in patients with filarial
tion and in vivo exposure. Bull Tech Gattefosse 104:61–69. chyluria. J Lipid Res 20:941–945.
Porter CJH and Charman WN (2001a) In vitro assessment of oral lipid based for- Rabinow B, Kipp J, Papadopoulos P, Wong J, Glosson J, Gass J, Sun CS, Wielgos T,
mulations. Adv Drug Deliv Rev 50 (Suppl 1):S127–S147. White R, and Cook C, et al. (2007) Itraconazole IV nanosuspension enhances effi-
Porter CJH and Charman WN (2001b) Intestinal lymphatic drug transport: an up- cacy through altered pharmacokinetics in the rat. Int J Pharm 339:251–260.
date. Adv Drug Deliv Rev 50:61–80. Rabinow BE (2004) Nanosuspensions in drug delivery. Nat Rev Drug Discov 3:
Porter CJH and Charman WN (2001c) Lipid-based formulations for oral adminis- 785–796.
tration: opportunities for bioavailability enhancement and lipoprotein targeting of Raghavan KS, Nemeth GA, Gray DB, and Hussain MA (1996) Solubility enhance-
lipophilic drugs. J Recept Signal Transduct Res 21:215–257. ment of a bisnaphthalimide tumoricidal agent, DMP 840, through complexation.
Porter CJH and Charman WN (2001d) Transport and absorption of drugs via the Pharm Dev Technol 1:231–238.
lymphatic system. Adv Drug Deliv Rev 50:1–2. Raghavan SL, Trividic A, Davis AF, and Hadgraft J (2001) Crystallization of hy-
Porter CJH, Kaukonen AM, Boyd BJ, Edwards GA, and Charman WN (2004a) drocortisone acetate: influence of polymers. Int J Pharm 212:213–221.
Susceptibility to lipase-mediated digestion reduces the oral bioavailability of Rajaraman G, Roberts MS, Hung D, Wang GQ, and Burczynski FJ (2005) Membrane
danazol after administration as a medium-chain lipid-based microemulsion for- binding proteins are the major determinants for the hepatocellular trans-
mulation. Pharm Res 21:1405–1412. membrane flux of long-chain fatty acids bound to albumin. Pharm Res 22:
Porter CJH, Kaukonen AM, Taillardat-Bertschinger A, Boyd BJ, O’Connor JM, 1793–1804.
Edwards GA, and Charman WN (2004b) Use of in vitro lipid digestion data to Rajewski RA and Stella VJ (1996) Pharmaceutical applications of cyclodextrins. 2. In
explain the in vivo performance of triglyceride-based oral lipid formulations of vivo drug delivery. J Pharm Sci 85:1142–1169.
poorly water-soluble drugs: studies with halofantrine. J Pharm Sci 93:1110–1121. Rama Prasad YV, Eaimtrakarn S, Ishida M, Kusawake Y, Tawa R, Yoshikawa Y,
Porter CJH, Trevaskis NL, and Charman WN (2007) Lipids and lipid-based for- Shibata N, and Takada K (2003) Evaluation of oral formulations of gentamicin
mulations: optimizing the oral delivery of lipophilic drugs. Nat Rev Drug Discov 6: containing labrasol in beagle dogs. Int J Pharm 268:13–21.
231–248. Rama Prasad YV, Minamimoto T, Yoshikawa Y, Shibata N, Mori S, Matsuura A,
Porter CJH, Wasan KM, and Constantinides P (2008) Lipid-based systems for the and Takada K (2004) In situ intestinal absorption studies on low molecular weight
enhanced delivery of poorly water soluble drugs. Adv Drug Deliv Rev 60:615–616. heparin in rats using labrasol as absorption enhancer. Int J Pharm 271:225–232.
Pouton CW (2000) Lipid formulations for oral administration of drugs: non- Ran YQ, Jain A, and Yalkowsky SH (2005) Solubilization and preformulation studies
emulsifying, self-emulsifying and ‘self-microemulsifying’ drug delivery systems. on PG-300995 (an anti-HIV drug). J Pharm Sci 94:297–303.
Eur J Pharm Sci 11 (Suppl 2):S93–S98. Randall K, Cheng S, and Kotchevar A. (2011) Evaluation of surfactants as solubi-
Pouton CW (2006) Formulation of poorly water-soluble drugs for oral administration: lizing agents in microsomal metabolism reactions with lipophilic substrates. In
physicochemical and physiological issues and the lipid formulation classification Vitro Cell Dev Biol Anim 47:631–639.
system. Eur J Pharm Sci 29:278–287. Rangel-Yagui CO, Pessoa A Jr, and Tavares LC (2005) Micellar solubilization of
Prajapati HN, Dalrymple DM, and Serajuddin ATM (2012) A comparative evaluation drugs. J Pharm Pharm Sci 8:147–165.
of mono-, di- and triglyceride of medium chain fatty acids by lipid/surfactant/water Rao VM, Haslam JL, and Stella VJ (2001) Controlled and complete release of a model
phase diagram, solubility determination and dispersion testing for application in poorly water-soluble drug, prednisolone, from hydroxypropyl methylcellulose ma-
pharmaceutical dosage form development. Pharm Res 29:285–305. trix tablets using (SBE)(7m)-beta-cyclodextrin as a solubilizing agent. J Pharm Sci
Pralhad T and Rajendrakumar K (2004) Study of freeze-dried quercetin-cyclodextrin 90:807–816.
binary systems by DSC, FT-IR, X-ray diffraction and SEM analysis. J Pharm Rao VM, Nerurkar M, Pinnamaneni S, Rinaldi F, and Raghavan K (2006) Co-
Biomed Anal 34:333–339. solubilization of poorly soluble drugs by micellization and complexation. Int J
Prasad YV, Puthli SP, Eaimtrakarn S, Ishida M, Yoshikawa Y, Shibata N, Pharm 319:98–106.
and Takada K (2003) Enhanced intestinal absorption of vancomycin with Labrasol Rao Z, Si L, Guan Y, Pan H, Qiu J, and Li G (2010) Inhibitive effect of cremophor
and D-alpha-tocopheryl PEG 1000 succinate in rats. Int J Pharm 250:181–190. RH40 or tween 80-based self-microemulsiflying drug delivery system on cyto-
Prestidge CA, Barnes TJ, Lau CH, Barnett C, Loni A, and Canham L (2007) chrome P450 3A enzymes in murine hepatocytes. J Huazhong Univ Sci Technolog
Mesoporous silicon: a platform for the delivery of therapeutics. Expert Opin Drug Med Sci 30:562–568.
Deliv 4:101–110. Rasenack N and Müller BW (2002) Dissolution rate enhancement by in situ
Preté PSC, Gomes K, Malheiros SVP, Meirelles NC, and de Paula E (2002) Solubi- micronization of poorly water-soluble drugs. Pharm Res 19:1894–1900.
lization of human erythrocyte membranes by non-ionic surfactants of the poly- Rasenack N and Müller BW (2004) Micron-size drug particles: common and novel
oxyethylene alkyl ethers series. Biophys Chem 97:45–54. micronization techniques. Pharm Dev Technol 9:1–13.
Prevosto D, Capaccioli S, Lucchesi M, Rolla PA, and Ngai KL (2009) Does the entropy Rawlinson CF, Williams AC, Timmins P, and Grimsey I (2007) Polymer-mediated
and volume dependence of the structural alpha-relaxation originate from the disruption of drug crystallinity. Int J Pharm 336:42–48.
Johari-Goldstein beta-relaxation? J Non-Cryst Solids 355:705–711. Ray A and Némethy G (1971) Effects of ionic protein denaturants on micelle for-
Prevosto D, Capaccioli S, Lucchesi M, Sharifi S, Kessairi K, and Ngai KL (2008) mation by nonionic detergents. J Am Chem Soc 93:6787–6793.
Relationship between structural and secondary relaxation in glass formers: Ratio Reddy MN, Rehana T, Ramakrishna S, Chowdhary KP, and Diwan PV (2004) Beta-
between glass transition temperature and activation energy. Philos Mag 88: cyclodextrin complexes of celecoxib: molecular-modeling, characterization, and
4063–4069. dissolution studies. AAPS PharmSci 6:E7.
Priestley RD, Ellison CJ, Broadbelt LJ, and Torkelson JM (2005) Structural re- Reddy LH, Sharma RK, Chuttani K, Mishra AK, and Murthy RSR (2005) Influence of
laxation of polymer glasses at surfaces, interfaces, and in between. Science 309: administration route on tumor uptake and biodistribution of etoposide loaded solid
456–459. lipid nanoparticles in Dalton’s lymphoma tumor bearing mice. J Control Release
Pudipeddi M and Serajuddin ATM (2005) Trends in solubility of polymorphs. 105:185-198.
J Pharm Sci 94:929–939. Redenti E, Szente L, and Szejtli J (2000) Drug/cyclodextrin/hydroxy acid multicom-
Pudipeddi M, Serajuddin ATM, Grant DJW, and Stahl PH (2002) Solubility and ponent systems: properties and pharmaceutical applications. J Pharm Sci 89:1–8.
dissolution of weak acids, bases and salts, in Handbook of Pharmaceutical Salts: Reed KW and Yalkowsky SH (1985) Lysis of human red blood cells in the presence of
Properties, Selection and Use (Stahl PH and Wermuth GH eds) pp 19–40, Wiley- various cosolvents. J Parenter Sci Technol 39:64–69.
VCH, Weinheim, Germany. Reed KW and Yalkowsky SH (1986) Lysis of human red blood cells in the presence of
Puri A, Loomis K, Smith B, Lee JH, Yavlovich A, Heldman E, and Blumenthal R various cosolvents. II. The effect of differing NaCl concentrations. J Parenter Sci
(2009) Lipid-based nanoparticles as pharmaceutical drug carriers: from concepts to Technol 40:88–94.
clinic. Crit Rev Ther Drug Carrier Syst 26:523–580. Reer O and Muller BW (1993) Investigation of the influence of cosolvents and sur-
Qi S and Craig DQM (2010) Detection of phase separation in hot melt extruded solid factants on the complexation of dexamethasone with hydroxypropyl-beta-
dispersion formulations. Am Pharm Rev 68–74. cyclodextrin by use of a simplex lattice design. Eur J Pharm Biopharm 39:105–111.
Qian F, Huang J, and Hussain MA (2010a) Drug-polymer solubility and miscibility: Rege BD, Kao JP, and Polli JE (2002) Effects of nonionic surfactants on membrane
stability consideration and practical challenges in amorphous solid dispersion de- transporters in Caco-2 cell monolayers. Eur J Pharm Sci 16:237–246.
velopment. J Pharm Sci 99:2941–2947. Regev R, Assaraf YG, and Eytan GD (1999) Membrane fluidization by ether, other
Qian F, Huang J, Zhu Q, Haddadin R, Gawel J, Garmise R, and Hussain M (2010b) Is anesthetics, and certain agents abolishes P-glycoprotein ATPase activity and
a distinctive single Tg a reliable indicator for the homogeneity of amorphous solid modulates efflux from multidrug-resistant cells. Eur J Biochem 259:18–24.
dispersion? Int J Pharm 395:232–235. Rekharsky M and Inoue Y (2000) Chiral recognition thermodynamics of beta-
Qian KK and Bogner RH (2011) Spontaneous crystalline-to-amorphous phase cyclodextrin: the thermodynamic origin of enantioselectivity and the enthalpy-
transformation of organic or medicinal compounds in the presence of porous media, entropy compensation effect. J Am Chem Soc 122:4418–4435.
part 1: thermodynamics of spontaneous amorphization. J Pharm Sci 100: Remenar JF, MacPhee JM, Larson BK, Tyagi VA, Ho JH, McIlroy DA, Hickey MB,
2801–2815. Shaw PB, and Almarsson O (2003a) Salt selection and simultaneous polymorphism
Qian KK and Bogner RH (2012) Application of mesoporous silicon dioxide and silicate assessment via high-throughput crystallization: the case of sertraline. Org Process
in oral amorphous drug delivery systems. J Pharm Sci 101:444–463. Res Dev 7:990–996.
Qian KK, Suib SL, and Bogner RH (2011) Spontaneous crystalline-to-amorphous Remenar JF, Morissette SL, Peterson ML, Moulton B, MacPhee JM, Guzmán HR,
phase transformation of organic or medicinal compounds in the presence of porous and Almarsson O (2003b) Crystal engineering of novel cocrystals of a triazole drug
media. Part 2. Amorphization capacity and mechanisms of interaction. J Pharm with 1,4-dicarboxylic acids. J Am Chem Soc 125:8456–8457.
Sci 100:4674–4686. Remenar JF, Peterson ML, Stephens PW, Zhang Z, Zimenkov Y, and Hickey MB
Qiao N, Li MZ, Schlindwein W, Malek N, Davies A, and Trappitt G (2011) Phar- (2007) Celecoxib:nicotinamide dissociation: using excipients to capture the cocrystal’s
maceutical cocrystals: an overview. Int J Pharm 419:1–11. potential. Mol Pharm 4:386–400.
Strategies for Low Drug Solubility 493
Ren X, Mao X, Cao L, Xue K, Si L, Qiu J, Schimmer AD, and Li G (2009) Nonionic Rumondor ACF, Ivanisevic I, Bates S, Alonzo DE, and Taylor LS (2009a) Evaluation
surfactants are strong inhibitors of cytochrome P450 3A biotransformation activity of drug-polymer miscibility in amorphous solid dispersion systems. Pharm Res 26:
in vitro and in vivo. Eur J Pharm Sci 36:401–411. 2523–2534.
Repka MA, Battu SK, Upadhye SB, Thumma S, Crowley MM, Zhang F, Martin C, Rumondor ACF, Stanford LA, and Taylor LS (2009b) Effects of polymer type and
and McGinity JW (2007) Pharmaceutical applications of hot-melt extrusion: Part storage relative humidity on the kinetics of felodipine crystallization from amor-
II. Drug Dev Ind Pharm 33:1043–1057. phous solid dispersions. Pharm Res 26:2599–2606.
Reppas C, Meyer JH, Sirois PJ, and Dressman JB (1991) Effect of hydrox- Rumondor ACF and Taylor LS (2010) Effect of polymer hygroscopicity on the phase
ypropylmethylcellulose on gastrointestinal transit and luminal viscosity in dogs. behavior of amorphous solid dispersions in the presence of moisture. Mol Pharm 7:
Gastroenterology 100:1217–1223. 477–490.
Reverchon E, De Marco I, and Della Porta G (2002) Rifampicin microparticles pro- Sachs-Barrable K, Lee SD, Wasan EK, Thornton SJ, and Wasan KM (2008) En-
duction by supercritical antisolvent precipitation. Int J Pharm 243:83–91. hancing drug absorption using lipids: a case study presenting the development and
Reverchon E, De Marco I, and Torino E (2007) Nanoparticles production by super- pharmacological evaluation of a novel lipid-based oral amphotericin B formulation
critical antisolvent precipitation: a general interpretation. J Supercrit Fluids 43: for the treatment of systemic fungal infections. Adv Drug Deliv Rev 60:692–701.
126–138. Sachs-Barrable K, Thamboo A, Lee SD, and Wasan KM (2007) Lipid excipients
Ribeiro A, Figueiras A, Santos D, and Veiga F (2008) Preparation and solid-state Peceol and Gelucire 44/14 decrease P-glycoprotein mediated efflux of rhodamine
characterization of inclusion complexes formed between miconazole and methyl- 123 partially due to modifying P-glycoprotein protein expression within Caco-2
beta-cyclodextrin. AAPS PharmSciTech 9:1102–1109. cells. J Pharm Pharm Sci 10:319–331.
Ribeiro Dos Santos I, Thies C, Richard J, Le Meurlay D, Gajan V, VandeVelde V, Saers ES and Craig DQM (1992) An investigation into the mechanisms of dissolution
and Benoit JP (2003) A supercritical fluid-based coating technology. 2: solubility of alkyl p-aminobenzoates from polyethylene-glycol solid dispersions. Int J Pharm
considerations. J Microencapsul 20:97–109. 83:211–219.
Ribeiro L, Carvalho RA, Ferreira DC, and Veiga FJB (2005a) Multicomponent Saers ES, Nystrom C, and Alden M (1993) Physicochemical aspects of drug release.
complex formation between vinpocetine, cyclodextrins, tartaric acid and water- 16. The effect of storage on drug dissolution from solid dispersions and the in-
soluble polymers monitored by NMR and solubility studies. Eur J Pharm Sci 24: fluence of cooling rate and incorporation of surfactant. Int J Pharm 90:105–118.
1–13. Sahoo NG, Kakran M, Shaal LA, Li L, Müller RH, Pal M, and Tan LP (2011) Prep-
Ribeiro L, Ferreira DC, and Veiga FJB (2005b) In vitro controlled release of aration and characterization of quercetin nanocrystals. J Pharm Sci 100:
vinpocetine-cyclodextrin-tartaric acid multicomponent complexes from HPMC 2379–2390.
swellable tablets. J Control Release 103:325–339. Salonen J, Laitinen L, Kaukonen AM, Tuura J, Björkqvist M, Heikkilä T, Vähä-
Ribeiro L, Loftsson T, Ferreira D, and Veiga F (2003a) Investigation and physico- Heikkilä K, Hirvonen J, and Lehto VP (2005) Mesoporous silicon microparticles for
chemical characterization of vinpocetine-sulfobutyl ether beta-cyclodextrin binary oral drug delivery: loading and release of five model drugs. J Control Release 108:
and ternary complexes. Chem Pharm Bull (Tokyo) 51:914–922. 362–374.
Ribeiro LSS, Ferreira DC, and Veiga FJB (2003b) Physicochemical investigation of Samaha MW and Gadalla MAF (1987) Solubilization of carbamazepin by different
the effects of water-soluble polymers on vinpocetine complexation with beta- classes of nonionic surfactants and a bile-salt. Drug Dev Ind Pharm 13:93–112.
cyclodextrin and its sulfobutyl ether derivative in solution and solid state. Eur J Sancin P, Caputo O, Cavallari C, Passerini N, Rodriguez L, Cini M, and Fini A (1999)
Pharm Sci 20:253–266. Effects of ultrasound-assisted compaction on Ketoprofen/Eudragit S100 mixtures.
Rigler MW, Honkanen RE, and Patton JS (1986) Visualization by freeze fracture, in Eur J Pharm Sci 7:207–213.
vitro and in vivo, of the products of fat digestion. J Lipid Res 27:836–857. Sanghvi R, Mogalian E, Machatha SG, Narazaki R, Karlage KL, Jain P, Tabibi SE,
Ring A, Pohl J, Völkl A, and Stremmel W (2002) Evidence for vesicles that mediate Glaze E, Myrdal PB, and Yalkowsky SH (2009) Preformulation and pharmaco-
long-chain fatty acid uptake by human microvascular endothelial cells. J Lipid Res kinetic studies on antalarmin: a novel stress inhibitor. J Pharm Sci 98:205–214.
Sarkar B, Lam S, and Alexandridis P (2010) Micellization of alkyl-propoxy-
43:2095–2104.
ethoxylate surfactants in water-polar organic solvent mixtures. Langmuir 26:
Rischin D, Webster LK, Millward MJ, Linahan BM, Toner GC, Woollett AM, Morton
10532–10540.
CG, and Bishop JF (1996) Cremophor pharmacokinetics in patients receiving 3-, 6-,
Sassene PJ, Knopp MM, Hesselkilde JZ, Koradia V, Larsen A, Rades T, and Müllertz
and 24-hour infusions of paclitaxel. J Natl Cancer Inst 88:1297–1301.
A (2010) Precipitation of a poorly soluble model drug during in vitro lipolysis:
Risovic V, Boyd M, Choo E, and Wasan KM (2003) Effects of lipid-based oral for-
characterization and dissolution of the precipitate. J Pharm Sci 99:4982–4991.
mulations on plasma and tissue amphotericin B concentrations and renal toxicity
Sathigari SK, Ober CA, Sanganwar GP, Gupta RB, and Babu RJ (2011) Single-step
in male rats. Antimicrob Agents Chemother 47:3339–3342.
preparation and deagglomeration of itraconazole microflakes by supercritical
Risovic V, Sachs-Barrable K, Boyd M, and Wasan KM (2004) Potential mechanisms
antisolvent method for dissolution enhancement. J Pharm Sci 100:2952–2965.
by which Peceol increases the gastrointestinal absorption of amphotericin B. Drug
Save T and Venkitachalam P (1992) Studies on solid dispersions of nifedipine. Drug
Dev Ind Pharm 30:767–774.
Dev Ind Pharm 18:1663–1679.
Rodgers J, Jones A, Gibaud S, Bradley B, McCabe C, Barrett MP, Gettinby G,
Savolainen J, Jarvinen K, Matilainen L, and Jarvinen T (1998) Improved dissolution
and Kennedy PGE (2011) Melarsoprol cyclodextrin inclusion complexes as prom- and bioavailability of phenytoin by sulfobutylether-beta-cyclodextrin ((SBE)(7m)-
ising oral candidates for the treatment of human African trypanosomiasis. PLoS beta-CD) and hydroxypropyl-beta-cyclodextrin (HP-beta-CD) complexation. Int J
Negl Trop Dis 5:e1308. Pharm 165:69–78.
Rodríguez-Hornedo N and Murphy D (1999) Significance of controlling crystalli- Schick JM (1966) Nonionic Surfactants, Marcel Dekker, New York.
zation mechanisms and kinetics in pharmaceutical systems. J Pharm Sci 88: Schinkel AH, Wagenaar E, Mol CA, and van Deemter L (1996) P-glycoprotein in the
651–660. blood-brain barrier of mice influences the brain penetration and pharmacological
Rodríguez-Spong B, Price CP, Jayasankar A, Matzger AJ, and Rodríguez-Hornedo N activity of many drugs. J Clin Invest 97:2517–2524.
(2004) General principles of pharmaceutical solid polymorphism: a supramolecular Schmidt LE and Dalhoff K (2002) Food-drug interactions. Drugs 62:1481–1502.
perspective. Adv Drug Deliv Rev 56:241–274. Schöler N, Krause K, Kayser O, Müller RH, Borner K, Hahn H, and Liesenfeld O
Rohrs BR, Thamann TJ, Gao P, Stelzer DJ, Bergren MS, and Chao RS (1999) Tablet (2001) Atovaquone nanosuspensions show excellent therapeutic effect in a new
dissolution affected by a moisture mediated solid-state interaction between drug murine model of reactivated toxoplasmosis. Antimicrob Agents Chemother 45:
and disintegrant. Pharm Res 16:1850–1856. 1771–1779.
Roos YH (1997) Frozen state transitions in relation to freeze drying. J Therm Anal Schöller-Gyüre M, Kakuda TN, Raoof A, De Smedt G, and Hoetelmans RMW (2009)
Calorim 48:535–544. Clinical pharmacokinetics and pharmacodynamics of etravirine. Clin Pharmaco-
Ross PD and Rekharsky MV (1996) Thermodynamics of hydrogen bond and hydro- kinet 48:561–574.
phobic interactions in cyclodextrin complexes. Biophys J 71:2144–2154. Schott H (1973) Effect of chain length in homologous series of anionic surfactants on
Rowe RC, Sheskey PJ, and Quinn ME eds (2009) Handbook of Pharmaceutical irritant action and toxicity. J Pharm Sci 62:341–343.
Excipients, Pharmaceutical Press, London. Schott H and Han SK (1975) Effect of inorganic additives on solutions of nonionic
Rowinsky EK, McGuire WP, Guarnieri T, Fisherman JS, Christian MC, surfactants II. J Pharm Sci 64:658–664.
and Donehower RC (1991) Cardiac disturbances during the administration of taxol. Schou M (1958) Lithium studies. 1. Toxicity. Acta Pharmacol Toxicol (Copenh) 15:
J Clin Oncol 9:1704–1712. 70–84.
Royall PG, Craig DQM, and Doherty C (1998) Characterisation of the glass transition Schrijvers D, Wanders J, Dirix L, Prove A, Vonck I, van Oosterom A, and Kaye S
of an amorphous drug using modulated DSC. Pharm Res 15:1117–1121. (1993) Coping with toxicities of docetaxel (Taxotere). Ann Oncol 4:610–611.
Rubin CE (1966) Electron microscopic studies of triglyceride absorption in man. Schuler US, Ehrsam M, Schneider A, Schmidt H, Deeg J, and Ehninger G (1998)
Gastroenterology 50:65–77. Pharmacokinetics of intravenous busulfan and evaluation of the bioavailability of
Rubino JT (1987) The effects of cosolvents on the action of pharmaceutical buffers. the oral formulation in conditioning for haematopoietic stem cell transplantation.
J Parenter Sci Technol 41:45–49. Bone Marrow Transplant 22:241–244.
Rubino JT(2002) Cosolvents and cosolvency, in Encylopedia of Pharmaceutical Schultheiss N and Newman A (2009) Pharmaceutical cocrystals and their physico-
Technology (Swarbrick J and Boylan JC, eds) Marcel Dekker, New York. chemical properties. Cryst Growth Des 9:2950–2967.
Rubino JT and Obeng EK (1991) Influence of solute structure on deviations from the Schulze JDR, Peters EE, Vickers AW, Staton JS, Coffin MD, Parsons GE, and Basit
log-linear solubility equation in propylene glycol:water mixtures. J Pharm Sci 80: AW (2005) Excipient effects on gastrointestinal transit and drug absorption in
479–483. beagle dogs. Int J Pharm 300:67–75.
Rubino JT and Thomas E (1990) Influence of solvent composition on the solubilities Schuurhuis GJ, Broxterman HJ, Pinedo HM, van Heijningen THM, van Kalken CK,
and solid-state properties of the sodium-salts of some drugs. Int J Pharm 65: Vermorken JB, Spoelstra EC, and Lankelma J (1990) The polyoxyethylene castor
141–145. oil Cremophor EL modifies multidrug resistance. Br J Cancer 62:591–594.
Rubino JT and Yalkowsky SH (1987) Cosolvency and deviations from log-linear Schwartz MA and Buckwalter FH (1962) Pharmaceutics of penicillin. J Pharm Sci
solubilization. Pharm Res 4:231–236. 51:1119–1128.
Ruble GR, Giardino OZ, Fossceco SL, Cosmatos D, Knapp RJ, and Barlow NJ (2006) Schwarz C and Mehnert W (1999) Solid lipid nanoparticles (SLN) for controlled drug
The effect of commonly used vehicles on canine hematology and clinical chemistry delivery. II. Drug incorporation and physicochemical characterization. J Micro-
values. J Am Assoc Lab Anim Sci 45:25–29. encapsul 16:205–213.
494 Williams et al.

Swenson ES and Curatolo WJ (1992) (C) Means to enhance penetration: (2) Intestinal Shah JC, Chen JR, and Chow D (1995) Preformulation study of etoposide. 2. In-
permeability enhancement for proteins, peptides and other polar drugs: mecha- creased solubility and dissolution rate by solid-solid dispersions. Int J Pharm 113:
nisms and potential toxicity. Adv Drug Deliv Rev 8:39–92. 103–111.
Seeballuck F, Ashford MB, and O’Driscoll CM (2003) The effects of pluronics block Shah JC, Sadhale Y, and Chilukuri DM (2001) Cubic phase gels as drug delivery
copolymers and Cremophor EL on intestinal lipoprotein processing and the po- systems. Adv Drug Deliv Rev 47:229–250.
tential link with P-glycoprotein in Caco-2 cells. Pharm Res 20:1085–1092. Shalel S, Streichman S, and Marmur A (2002) The mechanism of hemolysis by sur-
Seeballuck F, Lawless E, Ashford MB, and O’Driscoll CM (2004) Stimulation of factants: effect of solution composition. J Colloid Interface Sci 252:66–76.
triglyceride-rich lipoprotein secretion by polysorbate 80: in vitro and in vivo cor- Shamblin SL, Hancock BC, Dupuis Y, and Pikal MJ (2000) Interpretation of re-
relation using Caco-2 cells and a cannulated rat intestinal lymphatic model. Pharm laxation time constants for amorphous pharmaceutical systems. J Pharm Sci 89:
Res 21:2320–2326. 417–427.
Seedher N and Kanojia M (2009) Co-solvent solubilization of some poorly-soluble Shamblin SL, Tang XL, Chang LQ, Hancock BC, and Pikal MJ (1999) Character-
antidiabetic drugs. Pharm Dev Technol 14:185–192. ization of the time scales of molecular motion in pharmaceutically important
Seefeldt K, Miller J, Alvarez-Núñez F, and Rodríguez-Hornedo N (2007) Crystalli- glasses. J Phys Chem B 103:4113–4121.
zation pathways and kinetics of carbamazepine-nicotinamide cocrystals from the Shan N and Zaworotko MJ (2008) The role of cocrystals in pharmaceutical science.
amorphous state by in situ thermomicroscopy, spectroscopy, and calorimetry Drug Discov Today 13:440–446.
studies. J Pharm Sci 96:1147–1158. Shanker RM, Carola KV, Baltusis PJ, Brophy RT, and Hatfield TA (1994) Selection of
Seelig A and Gerebtzoff G (2006) Enhancement of drug absorption by noncharged appropriate salt form(s) for new drug candidates. Pharm Res 11:s-236.
detergents through membrane and P-glycoprotein binding. Expert Opin Drug Sharma P, Zujovic ZD, Bowmaker GA, Marshall AJ, Denny WA, and Garg S (2011)
Metab Toxicol 2:733–752. Evaluation of a crystalline nanosuspension: polymorphism, process induced
Segrave AM, Mager DE, Charman SA, Edwards GA, and Porter CJ (2004) Phar- transformation and in vivo studies. Int J Pharm 408:138–151.
macokinetics of recombinant human leukemia inhibitory factor in sheep. J Phar- Shayeganpour A, Jun AS, and Brocks DR (2005) Pharmacokinetics of amiodarone in
macol Exp Ther 309:1085–1092. hyperlipidemic and simulated high fat-meal rat models. Biopharm Drug Dispos 26:
Sek L, Boyd BJ, Charman WN, and Porter CJH (2006) Examination of the impact of 249–257.
a range of Pluronic surfactants on the in-vitro solubilisation behaviour and oral Shayeganpour A, Korashy H, Patel JP, El-Kadi AO, and Brocks DR (2008) The im-
bioavailability of lipidic formulations of atovaquone. J Pharm Pharmacol 58: pact of experimental hyperlipidemia on the distribution and metabolism of amio-
809–820. darone in rat. Int J Pharm 361:78–86.
Sek L, Porter CJH, Kaukonen AM, and Charman WN (2002) Evaluation of the in- Sheen PC, Khetarpal VK, Cariola CM, and Rowlings CE (1995) Formulation studies
vitro digestion profiles of long and medium chain glycerides and the phase be- of a poorly water-soluble drug in solid dispersions to improve bioavailability. Int J
haviour of their lipolytic products. J Pharm Pharmacol 54:29–41. Pharm 118:221–227.
Sekiguchi K and Obi N (1961) Studies on absorption of eutectic mixture. I. Com- Shen Q, Li W, Lin Y, Katsumi H, Okada N, Sakane T, Fujita T, and Yamamoto A
parison of behavior of eutectic mixture of sulfathiazole and that of ordinary sul- (2008) Modulating effect of polyethylene glycol on the intestinal transport and
fathiazole in man. Chem Pharm Bull (Tokyo) 9:866. absorption of prednisolone, methylprednisolone and quinidine in rats by in-vitro
Sekiguchi K, Obi N, and Ueda Y (1964) Studies on absorption of eutectic mixture. II. and in-situ absorption studies. J Pharm Pharmacol 60:1633–1641.
Absorption of fused conglomerates of chloramphenicol and urea in rabbits. Chem Shen Q, Lin YL, Handa T, Doi M, Sugie M, Wakayama K, Okada N, Fujita T,
Pharm Bull (Tokyo) 12:134–144. and Yamamoto A (2006) Modulation of intestinal P-glycoprotein function by poly-
Sekine M, Terashima H, Sasahara K, Nishimura K, Okada R, and Awazu S (1985) ethylene glycols and their derivatives by in vitro transport and in situ absorption
Improvement of bioavailability of poorly absorbed drugs. II. Effect of medium chain studies. Int J Pharm 313:49–56.
glyceride base on the intestinal absorption of cefmetazole sodium in rats and dogs. Shen S and Torkelson JM (1992) Miscibility and phase-separation in poly(methyl
J Pharmacobiodyn 8:286–295. methacrylate) poly(vinyl chloride) blends: study of thermodynamics by thermal-
Seo A, Holm P, Kristensen HG, and Schaefer T (2003) The preparation of agglom- analysis. Macromolecules 25:721–728.
erates containing solid dispersions of diazepam by melt agglomeration in a high Sheng JJ, Kasim NA, Chandrasekharan R, and Amidon GL (2006) Solubilization and
shear mixer. Int J Pharm 259:161–171. dissolution of insoluble weak acid, ketoprofen: effects of pH combined with sur-
Serajuddin ATM (1999) Solid dispersion of poorly water-soluble drugs: early factant. Eur J Pharm Sci 29:306–314.
promises, subsequent problems, and recent breakthroughs. J Pharm Sci 88: Shi Y and Li LC (2005) Current advances in sustained-release systems for parenteral
1058–1066. drug delivery. Expert Opin Drug Deliv 2:1039–1058.
Serajuddin ATM (2007) Salt formation to improve drug solubility. Adv Drug Deliv Shi Y, Porter W, Merdan T, and Li LC (2009) Recent advances in intravenous de-
Rev 59:603–616. livery of poorly water-soluble compounds. Expert Opin Drug Deliv 6:1261–1282.
Serajuddin ATM and Jarowski CI (1985a) Effect of diffusion layer pH and solubility Shiau YF (1981) Mechanisms of intestinal fat absorption. Am J Physiol 240:G1–G9.
on the dissolution rate of pharmaceutical acids and their sodium salts. II. Salicylic Shiau YF, Fernandez P, Jackson MJ, and McMonagle S (1985) Mechanisms main-
acid, theophylline, and benzoic acid. J Pharm Sci 74:148–154. taining a low-pH microclimate in the intestine. Am J Physiol 248:G608–G617.
Serajuddin ATM and Jarowski CI (1985b) Effect of diffusion layer pH and solubility Shimpi SL, Chauhan B, Mahadik KR, and Paradkar A (2005) Stabilization and im-
on the dissolution rate of pharmaceutical bases and their hydrochloride salts. I. proved in vivo performance of amorphous etoricoxib using Gelucire 50/13. Pharm
Phenazopyridine. J Pharm Sci 74:142–147. Res 22:1727–1734.
Serajuddin ATM and Pudipeddi M (2002) Salt-selection strategies, in Handbook of Shin SC and Kim J (2003) Physicochemical characterization of solid dispersion of
Pharmaceutical Salts: Properties, Selection and Use (Stahl PH and Wermuth GH, furosemide with TPGS. Int J Pharm 251:79–84.
eds) pp 135–160, Wiley-VCH, Weinheim. Shinde VR, Shelake MR, Shetty SS, Chavan-Patil AB, Pore YV, and Late SG (2008)
Serajuddin ATM, Sheen PC, and Augustine MA (1987) Common ion effect on solu- Enhanced solubility and dissolution rate of lamotrigine by inclusion complexation
bility and dissolution rate of the sodium salt of an organic acid. J Pharm Phar- and solid dispersion technique. J Pharm Pharmacol 60:1121–1129.
macol 39:587–591. Shinozaki H, Oguchi T, Suzuki S, Aoki K, Sako T, Morishita S, Tozuka Y, Moribe K,
Serajuddin ATM, Sheen PC, and Augustine MA (1990) Improved dissolution of and Yamamoto K (2006) Micronization and polymorphic conversion of tolbutamide
a poorly water-soluble drug from solid dispersions in polyethylene glycol: poly- and barbital by rapid expansion of supercritical solutions. Drug Dev Ind Pharm 32:
sorbate 80 mixtures. J Pharm Sci 79:463–464. 877–891.
Serajuddin ATM, Sheen PC, Mufson D, Bernstein DF, and Augustine MA (1986) Shiraki K, Takata N, Takano R, Hayashi Y, and Terada K (2008) Dissolution im-
Preformulation study of a poorly water-soluble drug, alpha-pentyl-3-(2- provement and the mechanism of the improvement from cocrystallization of poorly
quinolinylmethoxy)benzenemethanol: selection of the base for dosage form de- water-soluble compounds. Pharm Res 25:2581–2592.
sign. J Pharm Sci 75:492–496. Shono Y, Jantratid E, and Dressman JB (2011) Precipitation in the small intestine
Serajuddin ATM, Sheen PC, Mufson D, Bernstein DF, and Augustine MA (1988) may play a more important role in the in vivo performance of poorly soluble weak
Effect of vehicle amphiphilicity on the dissolution and bioavailability of a poorly bases in the fasted state: case example nelfinavir. Eur J Pharm Biopharm 79:
water-soluble drug from solid dispersions. J Pharm Sci 77:414–417. 349–356.
Serratoni M, Newton M, Booth S, and Clarke A (2007) Controlled drug release from Shono Y, Nishihara H, Matsuda Y, Furukawa S, Okada N, Fujita T, and Yamamoto A
pellets containing water-insoluble drugs dissolved in a self-emulsifying system. (2004) Modulation of intestinal P-glycoprotein function by cremophor EL and other
Eur J Pharm Biopharm 65:94–98. surfactants by an in vitro diffusion chamber method using the isolated rat in-
Sethia S and Squillante E (2002) Physicochemical characterization of solid dis- testinal membranes. J Pharm Sci 93:877–885.
persions of carbamazepine formulated by supercritical carbon dioxide and con- Sigfridsson K, Björkman JA, Skantze P, and Zachrisson H (2011a) Usefulness of
ventional solvent evaporation method. J Pharm Sci 91:1948–1957. a nanoparticle formulation to investigate some hemodynamic parameters of
Sethia S and Squillante E (2004) Solid dispersion of carbamazepine in PVP K30 by a poorly soluble compound. J Pharm Sci 100:2194–2202.
conventional solvent evaporation and supercritical methods. Int J Pharm 272:1–10. Sigfridsson K, Forssén S, Holländer P, Skantze U, and de Verdier J (2007) A for-
Sha X, Yan G, Wu Y, Li J, and Fang X (2005) Effect of self-microemulsifying drug mulation comparison, using a solution and different nanosuspensions of a poorly
delivery systems containing Labrasol on tight junctions in Caco-2 cells. Eur J soluble compound. Eur J Pharm Biopharm 67:540–547.
Pharm Sci 24:477–486. Sigfridsson K, Nordmark A, Theilig S, and Lindahl A (2011b) A formulation com-
Shackleford DM, Faassen WA, Houwing N, Lass H, Edwards GA, Porter CJH, parison between micro- and nanosuspensions: the importance of particle size for
and Charman WN (2003) Contribution of lymphatically transported testosterone absorption of a model compound, following repeated oral administration to rats
undecanoate to the systemic exposure of testosterone after oral administration of during early development. Drug Dev Ind Pharm 37:185–192.
two andriol formulations in conscious lymph duct-cannulated dogs. J Pharmacol Simha R and Boyer RF (1962) On a general relation involving the glass temperature
Exp Ther 306:925–933. and coefficients of expansion of polymers. J Chem Phys 37:1003–1007.
Shah AK and Agnihotri SA (2011) Recent advances and novel strategies in pre- Simonelli AP, Mehta SC, and Higuchi WI (1976) Dissolution rates of high energy
clinical formulation development: an overview. J Control Release 156:281–296. sulfathiazole-povidone coprecipitates. II. Characterization of form of drug con-
Shah AV and Serajuddin AT (2012) Development of solid self-emulsifying drug de- trolling its dissolution rate via solubility studies. J Pharm Sci 65:355–361.
livery system (SEDDS) I: Use of poloxamer 188 as both solidifying and emulsifying Singhal D and Curatolo W (2004) Drug polymorphism and dosage form design:
agent for lipids. Pharm Res 29:2817–2832. a practical perspective. Adv Drug Deliv Rev 56:335–347.
Strategies for Low Drug Solubility 495
Singla AK, Garg A, and Aggarwal D (2002) Paclitaxel and its formulations. Int J crystallization. Part 2. Analysis of 12 carboxylic acid co-crystals. J Pharm Sci
Pharm 235:179–192. 100:2734–2743.
Sinha S, Baboota S, Ali M, Kumar A, and Ali J (2009) Solid dispersion: an alternative Stanton MK, Tufekcic S, Morgan C, and Bak A (2009) Drug substance and former
technique for bioavailability enhancement of poorly soluble drugs. J Dispersion Sci structure property relationships in 15 diverse pharmaceutical co-crystals. Cryst
Technol 30:1458–1473. Growth Des 9:1344–1352.
Sinko PJ (2006) Martin’s Physical Pharmacy and Pharmaceutical Sciences, Stella VJ and He QR (2008) Cyclodextrins. Toxicol Pathol 36:30–42.
Lippincott Williams and Wilkins, Philadelphia. Stella VJ, Lee HK, and Thompson DO (1995) The effect of SBE4-beta-CD on IV
Sipahi AM, Oliveira HC, Vasconcelos KS, Castilho LN, Bettarello A, and Quintão EC methylprednisolone pharmacokinetics in rats: comparison to a cosolvent solution
(1989) Contribution of plasma protein and lipoproteins to intestinal lymph: com- and 2 water-soluble prodrugs. Int J Pharm 120:189–195.
parison of long-chain with medium-chain triglyceride duodenal infusion. Lym- Stella VJ and Nti-Addae KW (2007) Prodrug strategies to overcome poor water sol-
phology 22:13–19. ubility. Adv Drug Deliv Rev 59:677–694.
Sirois P (2007) Feasibility assessment and considerations for scaling initial prototype Stella VJ and Rajewski RA (1997) Cyclodextrins: their future in drug formulation and
lipid-based formulations to phase I/II clinical trial batches, in Oral Lipid-Based delivery. Pharm Res 14:556–567.
Formulations: Enhancing the Bioavailability of Poorly Water Soluble Drugs (Hauss Stella VJ, Rao VM, Zannou EA, and Zia V (1999) Mechanisms of drug release from
DJ ed) pp 63–78, Informa Healthcare, Inc., New York. cyclodextrin complexes. Adv Drug Deliv Rev 36:3–16.
Six K, Berghmans H, Leuner C, Dressman J, Van Werde K, Mullens J, Benoist L, Stevens DA (1999) Itraconazole in cyclodextrin solution. Pharmacotherapy 19:
Thimon M, Meublat L, and Verreck G, et al.. (2003a) Characterization of solid 603–611.
dispersions of itraconazole and hydroxypropylmethylcellulose prepared by melt Stevens JS, Byard SJ, and Schroeder SLM (2010) Salt or co-crystal? Determination of
extrusion, Part II. Pharm Res 20:1047–1054. protonation state by X-ray photoelectron spectroscopy (XPS). J Pharm Sci 99:
Six K, Daems T, de Hoon J, Van Hecken A, Depre M, Bouche MP, Prinsen P, Verreck 4453–4457.
G, Peeters J, and Brewster ME, et al. (2005) Clinical study of solid dispersions of Streng WH, Hsi SK, Helms PE, and Tan HGH (1984) General treatment of pH-
itraconazole prepared by hot-stage extrusion. Eur J Pharm Sci 24:179–186. solubility profiles of weak acids and bases and the effects of different acids on the
Six K, Murphy J, Weuts I, Craig DQM, Verreck G, Peeters J, Brewster M, and Van den solubility of a weak base. J Pharm Sci 73:1679–1684.
Mooter G (2003b) Identification of phase separation in solid dispersions of itraconazole Strickley RG (2004) Solubilizing excipients in oral and injectable formulations.
and Eudragit E100 using microthermal analysis. Pharm Res 20:135–138. Pharm Res 21:201–230.
Six K, Verreck G, Peeters J, Binnemans K, Berghmans H, Augustijns P, Kinget R, Strickley RG (2007) Currently marketed oral lipid-based dosage forms: enhancing the
and Van den Mooter G (2001) Investigation of thermal properties of glassy itraco- bioavailability of poorly water soluble drugs, in Oral Lipid-Based Formulations:
nazole: identification of a monotropic mesophase. Thermochim Acta 376:175–181. Enhancing the Bioavailability of Poorly Water Soluble Drugs (Hauss DJ ed) pp
Six K, Verreck G, Peeters J, Brewster M, and Van Den Mooter G (2004) Increased 1–31, Informa Healthcare, Inc., New York.
physical stability and improved dissolution properties of itraconazole, a class II Strickley RG and Oliyai R (2007) Solubilizing vehicles for oral formulation de-
drug, by solid dispersions that combine fast- and slow-dissolving polymers. velopment, in Solvent Systems and Their Selection in Pharmaceutics and Bio-
J Pharm Sci 93:124–131. pharmaceutics (Augustijns P and Brewster M eds) pp 257–308, Springer, New York.
Sjostrom B and Bergenstahl B (1992) Preparation of submicron drug particles in Su CS, Tang M, and Chen YP (2009) Micronization of nabumetone using the rapid
lecithin-stabilized O/W emulsions. 1. Model studies of the precipitation cholesteryl expansion of supercritical solution (RESS) process. J Supercrit Fluids 50:69–76.
acetate. Int J Pharm 88:53–62. Subramaniam B, Rajewski RA, and Snavely K (1997) Pharmaceutical processing
Slack JD, Kanke M, Simmons GH, and DeLuca PP (1981) Acute hemodynamic effects with supercritical carbon dioxide. J Pharm Sci 86:885–890.
and blood pool kinetics of polystyrene microspheres following intravenous admin- Sugano K (2008) Theoretical comparison of hydrodynamic diffusion layer models
istration. J Pharm Sci 70:660–664. used for dissolution simulation in drug discovery and development. Int J Pharm
Smart T, Lomas H, Massignani M, Flores-Merino MV, Perez LR, and Battaglia G 363:73–77.
(2008) Block copolymer nanostructures. Nano Today 3:38–46. Suhonen P, Järvinen T, Lehmussaari K, Reunamäki T, and Urtti A (1995) Ocular
Smith AJ, Kavuru P, Wojtas L, Zaworotko MJ, and Shytle RD (2011) Cocrystals of absorption and irritation of pilocarpine prodrug is modified with buffer, polymer,
quercetin with improved solubility and oral bioavailability. Mol Pharm 8: and cyclodextrin in the eyedrop. Pharm Res 12:529–533.
1867–1876. Summers MP and Enever RP (1976) Preparation and properties of solid dispersion
Smith DA, Di L, and Kerns EH (2010) The effect of plasma protein binding on in vivo system containing citric acid and primidone. J Pharm Sci 65:1613–1617.
efficacy: misconceptions in drug discovery. Nat Rev Drug Discov 9:929–939. Sun W, Mao SR, Shi Y, Li LC, and Fang L (2011) Nanonization of itraconazole by
Smith JS, Macrae RJ, and Snowden MJ (2005) Effect of SBE7-beta-cyclodextrin high pressure homogenization: stabilizer optimization and effect of particle size on
complexation on carbamazepine release from sustained release beads. Eur J oral absorption. J Pharm Sci 100:3365–3373.
Pharm Biopharm 60:73–80. Sun Y, Tao J, Zhang GGZ, and Yu LA (2010) Solubilities of crystalline drugs in
Snodin DJ (2006) Residues of genotoxic alkyl mesylates in mesylate salt drug sub- polymers: an improved analytical method and comparison of solubilities of in-
stances: real or imaginary problems? Regul Toxicol Pharmacol 45:79–90. domethacin and nifedipine in PVP, PVP/VA, and PVAc. J Pharm Sci 99:
Söderlind E, Wollbratt M, and von Corswant C (2003) The usefulness of sugar sur- 4023–4031.
factants as solubilizing agents in parenteral formulations. Int J Pharm 252:61–71. Sunesen VH, Vedelsdal R, Kristensen HG, Christrup L, and Mullertz A (2005) Effect
Sokol RJ, Johnson KE, Karrer FM, Narkewicz MR, Smith D, and Kam I (1991) of liquid volume and food intake on the absolute bioavailability of danazol, a poorly
Improvement of cyclosporin absorption in children after liver transplantation by soluble drug. Eur J Pharmaceutical Sci 24:297–303.
means of water-soluble vitamin E. Lancet 338:212–214. Suresh G, Manjunath K, Venkateswarlu V, and Satyanarayana V (2007) Prepara-
Sollohub K and Cal K (2010) Spray drying technique. II. Current applications in tion, characterization, and in vitro and in vivo evaluation of lovastatin solid lipid
pharmaceutical technology. J Pharm Sci 99:587–597. nanoparticles. AAPS PharmSciTech 8:E162–E170.
Sotthivirat S, Haslam JL, and Stella VJ (2007) Controlled porosity-osmotic pump Suzuki H and Sunada H (1998) Influence of water-soluble polymers on the dissolu-
pellets of a poorly water-soluble drug using sulfobutylether-beta-cyclodextrin, tion of nifedipine solid dispersions with combined carriers. Chem Pharm Bull
(SBE) 7M-beta-CD, as a solubilizing and osmotic agent. J Pharm Sci 96: (Tokyo) 46:482–487.
2364–2374. Szatmari I and Vargay Z (1988) Pharmacokinetics of dimethyl-beta-cyclodextrin in
Southard MZ, Green DW, Stella VJ, and Himmelstein KJ (1992) Dissolution of ion- rats, in Proceedings of the 4th International Symposium on Cyclodextrins (Duchene
izable drugs into unbuffered solution: a comprehensive model for mass transport D, ed) pp 407–413; 1988 April 20–22; Munich, Germany; Kluwer Academic Pub-
and reaction in the rotating disk geometry. Pharm Res 9:58–69. lishers, Norwell, MA.
Sparreboom A, van Zuylen L, Brouwer E, Loos WJ, de Bruijn P, Gelderblom H, Pillay Szejtli J (1994) Medicinal applications of cyclodextrins. Med Res Rev 14:353–386.
M, Nooter K, Stoter G, and Verweij J (1999) Cremophor EL-mediated alteration of Szejtli J (1998) Introduction and general overview of cyclodextrin chemistry. Chem
paclitaxel distribution in human blood: clinical pharmacokinetic implications. Rev 98:1743–1754.
Cancer Res 59:1454–1457. Szente L and Szejtli J (1999) Highly soluble cyclodextrin derivatives: chemistry,
Spoelstra EC, Dekker H, Schuurhuis GJ, Broxterman HJ, and Lankelma J (1991) P- properties, and trends in development. Adv Drug Deliv Rev 36:17–28.
glycoprotein drug efflux pump involved in the mechanisms of intrinsic drug re- Tadros TF (2005) Applied Surfactants: Principles and Applications, Wiley-VCH, New
sistance in various colon cancer cell lines: evidence for a saturation of active York.
daunorubicin transport. Biochem Pharmacol 41:349–359. Taillardat A, Diederich A, Sutter B, Kalb O, and Cuiné JF (2007) Use of in vitro
Staggers JE, Hernell O, Stafford RJ, and Carey MC (1990) Physical-chemical be- dispersion and digestion tests to explain the in vivo performance of two lipid- based
havior of dietary and biliary lipids during intestinal digestion and absorption. 1. oral drug delivery systems in man. Am Pharm Rev 10:40–46.
Phase behavior and aggregation states of model lipid systems patterned after Tajarobi F, Larsson A, Matic H, and Abrahmsén-Alami S (2011) The influence of
aqueous duodenal contents of healthy adult human beings. Biochemistry 29: crystallization inhibition of HPMC and HPMCAS on model substance dissolu-
2028–2040. tion and release in swellable matrix tablets. Eur J Pharm Biopharm 78:
Stahl PH (2003) Preparation of water-soluble compounds through salt formation, in 125–133.
The Practice of Medicinal Chemistry, 2nd Edition (Wermuth CG ed) pp 601–615, Takagi T, Ramachandran C, Bermejo M, Yamashita S, Yu LX, and Amidon GL (2006)
London, Elsevier. A provisional biopharmaceutical classification of the top 200 oral drug products in
Stahl PH and Wermuth GH (2002) Handbook of Pharmaceutical Salts: Properties, the United States, Great Britain, Spain, and Japan. Mol Pharm 3:631–643.
Selection and Use, Wiley-VCH, Weinheim. Takeuchi H, Nagira S, Yamamoto H, and Kawashima Y (2005) Solid dispersion
Stanton MK and Bak A (2008) Physicochemical properties of pharmaceutical co- particles of amorphous indomethacin with fine porous silica particles by using
crystals: A case study of ten AMG 517 co-crystals. Cryst Growth Des 8:3856–3862. spray-drying method. Int J Pharm 293:155–164.
Stanton MK, Kelly RC, Colletti A, Kiang YH, Langley M, Munson EJ, Peterson ML, Takisawa N, Shirahama K, and Tanaka I (1993) Interactions of amphiphilic drugs
Roberts J, and Wells M (2010) Improved pharmacokinetics of AMG 517 through co- with alpha-cyclodextrin, beta-cyclodextrin, and gamma-cyclodextrin. Colloid Polym
crystallization. Part 1: comparison of two acids with corresponding amide co- Sci 271:499–506.
crystals. J Pharm Sci 99:3769–3778. Tan A, Davey AK, and Prestidge CA (2011) Silica-lipid hybrid (SLH) versus non-lipid
Stanton MK, Kelly RC, Colletti A, Langley M, Munson EJ, Peterson ML, Roberts J, formulations for optimising the dose-dependent oral absorption of celecoxib. Pharm
and Wells M (2011) Improved pharmacokinetics of AMG 517 through co- Res 28:2273–2287.
496 Williams et al.

Tan A, Martin A, Nguyen T-H, Boyd BJ, and Prestidge CA (2012) Hybrid nano- alcohols: structural investigations by small-angle X-ray scattering and dynamic
materials that mimic the food effect: controlling enzymatic digestion for enhanced light scattering. J Colloid Interface Sci 294:194–211.
oral drug absorption. Angew Chem Int Ed Engl 51:5475–5479. Tong P and Zografi G (1999) Solid-state characteristics of amorphous sodium in-
Tan A, Simovic S, Davey AK, Rades T, and Prestidge CA (2009) Silica-lipid hybrid domethacin relative to its free acid. Pharm Res 16:1186–1192.
(SLH) microcapsules: a novel oral delivery system for poorly soluble drugs. J Tong P and Zografi G (2001) A study of amorphous molecular dispersions of in-
Control Release 134:62–70. domethacin and its sodium salt. J Pharm Sci 90:1991–2004.
Tan XY and Lindenbaum S (1991) Studies on complexation between beta- Tong WQ, Lach JL, Chin TF, and Guillory JK (1991) Structural effects on the binding
cyclodextrin and bile-salts. Int J Pharm 74:127–135. of amine drugs with the diphenylmethyl functionality to cyclodextrins. I. A mi-
Tanford C (1974) Thermodynamics of micelle formation: prediction of micelle size and crocalorimetric study. Pharm Res 8:951–957.
size distribution. Proc Natl Acad Sci USA 71:1811–1815. Tong WQ and Whitesell G (1998) In situ salt screening: a useful technique for dis-
Tanford C (1980) The Hydrophobic Rffect: Formation of Micelles and Biological covery support and preformulation studies. Pharm Dev Technol 3:215–223.
Membranes, Wiley, New York. Tønnesen HH, Másson M, and Loftsson T (2002) Studies of curcumin and curcumi-
Tang B, Cheng G, Gu JC, and Xu CH (2008) Development of solid self-emulsifying noids. XXVII. Cyclodextrin complexation: solubility, chemical and photochemical
drug delivery systems: preparation techniques and dosage forms. Drug Discov stability. Int J Pharm 244:127–135.
Today 13:606–612. Tønsberg H, Holm R, Mu H, Boll JB, Jacobsen J, and Müllertz A (2011) Effect of bile
Taniguchi T, Yamamoto K, and Kobayashi T (1998) Precipitate formed by thio- on the oral absorption of halofantrine in polyethylene glycol 400 and polysorbate
pentone and vecuronium causes pulmonary embolism. Can J Anaesth 45:347–351. 80 formulations dosed to bile duct cannulated rats. J Pharm Pharmacol 63:
Tanno F, Nishiyama Y, Kokubo H, and Obara S (2004) Evaluation of hypromellose 817–824.
acetate succinate (HPMCAS) as a carrier in solid dispersions. Drug Dev Ind Pharm Torosian G, Finger KF, and Stewart RB (1973) Hazards of bromides in proprietary
30:9–17. medication. Am J Hosp Pharm 30:716–718.
Tantishaiyakul V, Kaewnopparat N, and Ingkatawornwong S (1999) Properties of Towler CS, Li TL, Wikström H, Remick DM, Sanchez-Felix MV, and Taylor LS (2008)
solid dispersions of piroxicam in polyvinylpyrrolidone. Int J Pharm 181:143–151. An investigation into the influence of counterion on the properties of some amor-
Tao T, Zhao Y, Wu JJ, and Zhou BY (2009) Preparation and evaluation of itraco- phous organic salts. Mol Pharm 5:946–955.
nazole dihydrochloride for the solubility and dissolution rate enhancement. Int J Tozuka Y, Fujito T, Moribe K, and Yamamoto K (2006) Ibuprofen-cyclodextrin in-
Pharm 367:109–114. clusion complex formation using supercritical carbon dioxide. J Incl Phenom
Tarsa PB, Towler CS, Woollam G, and Berghausen J (2010) The influence of aqueous Macrocycl Chem 56:33–37.
content in small scale salt screening: improving hit rate for weakly basic, low Trägner D and Csordas A (1987) Biphasic interaction of Triton detergents with the
solubility drugs. Eur J Pharm Sci 41:23–30. erythrocyte membrane. Biochem J 244:605–609.
Tavares LC (2007) Breakage of single particles: quasi-static, in Handbook of Powder Trask AV, Motherwell WDS, and Jones W (2004) Solvent-drop grinding: green
Technology (Williams JC and Allen T eds) pp 3–68, Elsevier, New York. polymorph control of cocrystallisation. Chem Commun (Camb) 7:890–891.
Taylor LS and Zografi G (1997) Spectroscopic characterization of interactions be- Treiner C (1994) The thermodynamics of micellar solubilization of neutral solutes in
tween PVP and indomethacin in amorphous molecular dispersions. Pharm Res 14: aqueous binary surfactant systems. Chem Soc Rev 23:349–356.
1691–1698. Trevaskis NL, Charman WN, and Porter CJH (2008) Lipid-based delivery systems
Tayrouz Y, Ding R, Burhenne J, Riedel KD, Weiss J, Hoppe-Tichy T, Haefeli WE, and intestinal lymphatic drug transport: a mechanistic update. Adv Drug Deliv Rev
and Mikus G (2003) Pharmacokinetic and pharmaceutic interaction between di- 60:702–716.
goxin and Cremophor RH40. Clin Pharmacol Ther 73:397–405. Trevaskis NL, McEvoy CL, McIntosh MP, Edwards GA, Shanker RM, Charman WN,
Teasdale A, Eyley SC, Delaney E, Jacq K, Taylor-Worth K, Lipczynski A, Reif V, and Porter CJH (2010a) The role of the intestinal lymphatics in the absorption of
Elder DP, Facchine KL, and Golec S, et al. (2009) Mechanism and processing two highly lipophilic cholesterol ester transfer protein inhibitors (CP524,515 and
parameters affecting the formation of methyl methanesulfonate from methanol CP532,623). Pharm Res 27:878–893.
and methanesulfonic acid: an illustrative example for sulfonate ester impurity Trevaskis NL, Porter CJH, and Charman WN (2006) An examination of the interplay
formation. Org Process Res Dev 13:429–433. between enterocyte-based metabolism and lymphatic drug transport in the rat.
ten Tije AJ, Verweij J, Loos WJ, and Sparreboom A (2003) Pharmacological effects of Drug Metab Dispos 34:729–733.
formulation vehicles: implications for cancer chemotherapy. Clin Pharmacokinet Trevaskis NL, Shackleford DM, Charman WN, Edwards GA, Gardin A, Appel-
42:665–685. Dingemanse S, Kretz O, Galli B, and Porter CJH (2009) Intestinal lymphatic
Thakur R and Gupta RB (2006) Formation of phenytoin nanoparticles using rapid transport enhances the post-prandial oral bioavailability of a novel cannabinoid
expansion of supercritical solution with solid cosolvent (RESS-SC) process. Int J receptor agonist via avoidance of first-pass metabolism. Pharm Res 26:1486–1495.
Pharm 308:190–199. Trevaskis NL, Shanker RM, Charman WN, and Porter CJH (2010b) The mechanism
Thayyil MS, Capaccioli S, Prevosto D, and Ngai KL (2008) Is the Johari-Goldstein of lymphatic access of two cholesteryl ester transfer protein inhibitors (CP524,515
beta-relaxation universal? Philos Mag 88:4007–4013. and CP532,623) and evaluation of their impact on lymph lipoprotein profiles.
The Merck Index (2006) 14th ed. Merck & Co. Inc., Whitehouse Station, NJ. Pharm Res 27:1949–1964.
Thies C, Ribeiro Dos Santos I, Richard J, VandeVelde V, Rolland H, and Benoit JP Tso P, Karlstad MD, Bistrian BR, and DeMichele SJ (1995) Intestinal digestion,
(2003) A supercritical fluid-based coating technology. 1. Process considerations. absorption, and transport of structured triglycerides and cholesterol in rats. Am J
J Microencapsul 20:87–96. Physiol 268:G568–G577.
Thomas E and Rubino J (1996) Solubility, melting point and salting-out relationships Tsui HW, Wang JH, Hsu YH, and Chen LJ (2010) Study of heat of micellization and
in a group of secondary amine hydrochloride salts. Int J Pharm 130:179–185. phase separation for Pluronic aqueous solutions by using a high sensitivity dif-
Thomas N, Holm R, Müllertz A, and Rades T (2012) In vitro and in vivo performance ferential scanning calorimetry. Colloid Polym Sci 288:1687–1696.
of novel supersaturated self-nanoemulsifying drug delivery systems (super- Turk M and Bolten D (2010) Formation of submicron poorly water-soluble drugs by
SNEDDS). J Control Release 160:25–32. rapid expansion of supercritical solution (RESS): results for naproxen. J Supercrit
Thombre AG, Herbig SM, and Alderman JA (2011) Improved ziprasidone for- Fluids 55:778–785.
mulations with enhanced bioavailability in the fasted state and a reduced food Turk M, Hils P, Helfgen B, Schaber K, Martin HJ, and Wahl MA (2002) Micron-
effect. Pharm Res 28:3159–3170. ization of pharmaceutical substances by the rapid expansion of supercritical sol-
Thombre AG, Shah JC, Sagawa K, and Caldwell WB (2012) In vitro and in vivo utions (RESS): a promising method to improve bioavailability of poorly soluble
characterization of amorphous, nanocrystalline, and crystalline ziprasidone for- pharmaceutical agents. J Supercrit Fluids 22:75–84.
mulations. Int J Pharm 428:8–17. Turnbull D (1969) Under what conditions can a glass be formed. Contemp Phys 10:473.
Thompson DO (1997) Cyclodextrins: enabling excipients: their present and future use Turnbull D and Fisher JC (1949) Rate of nucleation in condensed systems. J Chem
in pharmaceuticals. Crit Rev Ther Drug Carrier Syst 14:1–104. Phys 17:71–73.
Thuahnai ST, Lund-Katz S, Williams DL, and Phillips MC (2001) Scavenger receptor Tytgat GN, Rubin CE, and Saunders DR (1971) Synthesis and transport of lipopro-
class B, type I-mediated uptake of various lipids into cells. Influence of the nature tein particles by intestinal absorptive cells in man. J Clin Invest 50:2065–2078.
of the donor particle interaction with the receptor. J Biol Chem 276:43801–43808. Uekama K (2002) Recent aspects of pharmaceutical application of cyclodextrins. J
Tinwalla AY, Hoesterey BL, Xiang TX, Lim K, and Anderson BD (1993) Solubiliza- Incl Phenom Macrocycl Chem 44:3–7.
tion of thiazolobenzimidazole using a combination of pH adjustment and com- Uekama K (2004) Design and evaluation of cyclodextrin-based drug formulation.
plexation with 2-hydroxypropyl-beta-cyclodextrin. Pharm Res 10:1136–1143. Chem Pharm Bull (Tokyo) 52:900–915.
Tokumura T, Nanba M, Tsushima Y, Tatsuishi K, Kayano M, Machida Y, and Nagai Uekama K, Fujinaga T, Hirayama F, Otagiri M, and Yamasaki M (1982) Inclusion
T (1986) Enhancement of bioavailability of cinnarizine from its beta-cyclodextrin complexations of steroid-hormones with cyclodextrins in water and in solid-phase.
complex on oral administration with DL-phenylalanine as a competing agent. J Int J Pharm 10:1–15.
Pharm Sci 75:391–394. Uekama K, Hirayama F, and Irie T (1998) Cyclodextrin drug carrier systems. Chem
Tombari E, Ferrari C, Johari GP, and Shanker RM (2008a) Calorimetric relaxation in Rev 98:2045–2076.
pharmaceutical molecular glasses and its utility in understanding their stability Uekama K, Matsubara K, Abe K, Horiuchi Y, Hirayama F, and Suzuki N (1990)
against crystallization. J Phys Chem B 112:10806–10814. Design and in vitro evaluation of slow-release dosage form of piretanide: utility of
Tombari E, Ferrari C, Salvetti G, and Johari GP (2008b) Vibrational and configu- beta-cyclodextrin:cellulose derivative combination as a modified-release drug car-
rational specific heats during isothermal structural relaxation of a glass to equi- rier. J Pharm Sci 79:244–248.
librium liquid. Phys Rev B 78:144203-1–144203-8. Uekama K, Narisawa S, Hirayama F, and Otagiri M (1983) Improvement of disso-
Tomita M, Sawada T, Ogawa T, Ouchi H, Hayashi M, and Awazu S (1992) Differences lution and absorption characteristics of benzodiazepines by cyclodextrin complex-
in the enhancing effects of sodium caprate on colonic and jejunal drug absorption. ation. Int J Pharm 16:327–338.
Pharm Res 9:648–653. Unowsky J, Behl CR, Beskid G, Sattler J, Halpern J, and Cleeland R (1988) Effect of
Tommasini S, Calabrò ML, Raneri D, Ficarra P, and Ficarra R (2004) Combined medium chain glycerides on enteral and rectal absorption of beta-lactam and
effect of pH and polysorbates with cyclodextrins on solubilization of naringenin. J aminoglycoside antibiotics. Chemotherapy 34:272–276.
Pharm Biomed Anal 36:327–333. Usach I and Peris JE (2011) Bioavailability of nevirapine in rats after oral and
Tomsic M, Bester-Rogac M, Jamnik A, Kunz W, Touraud D, Bergmann A, and Glatter subcutaneous administration, in vivo absorption from gastrointestinal segments
O (2006) Ternary systems of nonionic surfactant Brij 35, water and various simple and effect of bile on its absorption from duodenum. Int J Pharm 419:186–191.
Strategies for Low Drug Solubility 497
Usui F, Maeda K, Kusai A, Nishimura K, and Yamamoto K (1997) Inhibitory effects nonstoichiometric cocrystals off a phosphodiesterase-IV inhibitor and L-tartaric
of water-soluble polymers on precipitation of RS-8359. Int J Pharm 154:59–66. acid. Cryst Growth Des 6:690–700.
Vaddi HK, Khan MA, and Reddy IK (2007) Ocular, nasal and otic drug delivery, in Varma MV and Panchagnula R (2005) Enhanced oral paclitaxel absorption with
Gibaldi’s Drug Delivery Systems in Pharmaceutical Care (Gibaldi M, Desai A, vitamin E-TPGS: effect on solubility and permeability in vitro, in situ and in vivo.
and Lee M eds) pp 59–78, Bethesda, MD, American Society of Health-System Eur J Pharmaceutical Sci 25:445–453.
Pharmacists. Vasanthavada M and Serajuddin ATM (2007) Lipid-based self-emulsifying solid dis-
Valero M, Pérez-Revuelta BI, and Rodríguez LJ (2003) Effect of PVP K-25 on the persions, in Oral Lipid-Based Formulations: Enhancing the Bioavailability of Poorly
formation of the naproxen:beta-ciclodextrin complex. Int J Pharm 253:97–110. Water-Soluble Drugs (Hauss DJ ed) pp 149–184, Informa Healthcare, New York.
Valizadeh H, Nokhodchi A, Qarakhani N, Zakeri-Milani P, Azarmi S, Hassanzadeh Vasanthavada M, Tong WQ, Joshi Y, and Kislalioglu MS (2004) Phase behavior of
D, and Löbenberg R (2004) Physicochemical characterization of solid dispersions of amorphous molecular dispersions. I. Determination of the degree and mechanism
indomethacin with PEG 6000, Myrj 52, lactose, sorbitol, dextrin, and Eudragit of solid solubility. Pharm Res 21:1598–1606.
E100. Drug Dev Ind Pharm 30:303–317. Vasanthavada M, Tong WQ, Joshi Y, and Kislalioglu MS (2005) Phase behavior of
van Asperen J, Schinkel AH, Beijnen JH, Nooijen WJ, Borst P, and van Tellingen O amorphous molecular dispersions. II. Role of hydrogen bonding in solid solubility
(1996) Altered pharmacokinetics of vinblastine in Mdr1a P-glycoprotein-deficient and phase separation kinetics. Pharm Res 22:440–448.
mice. J Natl Cancer Inst 88:994–999. Vasconcelos T, Sarmento B, and Costa P (2007) Solid dispersions as strategy to
van den Bosch H, Postema NM, de Haas GH, and van Deenen LL (1965) On the improve oral bioavailability of poor water soluble drugs. Drug Discov Today 12:
positional specificity of phospholipase A from pancreas. Biochim Biophys Acta 98: 1068–1075.
657–659. Vehring R (2008) Pharmaceutical particle engineering via spray drying. Pharm Res
Van den Mooter G, Augustijns P, Blaton N, and Kinget R (1998) Physico-chemical 25:999–1022.
characterization of solid dispersions of temazepam with polyethylene glycol 6000 Veiga F, Fernandes C, and Maincent P (2001) Influence of the preparation method on
and PVP K30. Int J Pharm 164:67–80. the physicochemical properties of tolbutamide/cyclodextrin binary systems. Drug
Vandecruys R, Peeters J, Verreck G, and Brewster ME (2007) Use of a screening Dev Ind Pharm 27:523–532.
method to determine excipients which optimize the extent and stability of super- Veiga F, Teixeira-Dias JJC, Kedzierewicz F, Sousa A, and Maincent P (1996) In-
saturated drug solutions and application of this system to solid formulation design. clusion complexation of tolbutamide with beta-cyclodextrin and hydroxypropyl-
Int J Pharm 342:168–175. beta-cyclodextrin. Int J Pharm 129:63–71.
Van den Mooter G, Wuyts M, Blaton N, Busson R, Grobet P, Augustijns P, and Kinget Veiga MD and Ahsan F (1998) Solubility study of tolbutamide in monocomponent and
R (2001) Physical stabilisation of amorphous ketoconazole in solid dispersions with dicomponent solutions of water. Int J Pharm 160:43–49.
polyvinylpyrrolidone K25. Eur J Pharm Sci 12:261–269. Veiga MD, Escobar C, and Bernad MJ (1993) Dissolution behavior of drugs from
Van Drooge DJ, Hinrichs WLJ, and Frijlink HW (2004) Incorporation of lipophilic binary and ternary-systems. Int J Pharm 93:215–220.
drugs in sugar glasses by lyophilization using a mixture of water and tertiary butyl Vekilov PG (2010) Nucleation. Cryst Growth Des 10:5007–5019.
alcohol as solvent. J Pharm Sci 93:713–725. Venkatesh S, Li JM, Xu YH, Vishnuvajjala R, and Anderson BD (1996) Intrinsic
van Drooge DJ, Hinrichs WLJ, Visser MR, and Frijlink HW (2006) Characterization solubility estimation and pH-solubility behavior of cosalane (NSC 658586), an
of the molecular distribution of drugs in glassy solid dispersions at the nano-meter extremely hydrophobic diprotic acid. Pharm Res 13:1453–1459.
scale, using differential scanning calorimetry and gravimetric water vapour sorp- Vennerstrom JL, Arbe-Barnes S, Brun R, Charman SA, Chiu FC, Chollet J, Dong Y,
tion techniques. Int J Pharm 310:220–229. Dorn A, Hunziker D, and Matile H, et al. (2004) Identification of an antimalarial
Van Eerdenbrugh B, Alonzo DE, and Taylor LS (2011) Influence of particle size on the synthetic trioxolane drug development candidate. Nature 430:900–904.
ultraviolet spectrum of particulate-containing solutions: implications for in-situ Verbeeck RK, Kanfer I, and Walker RB (2006) Generic substitution: the use of me-
concentration monitoring using UV/Vis fiber-optic probes. Pharm Res 28:1643–1652. dicinal products containing different salts and implications for safety and efficacy.
Van Eerdenbrugh B, Baird JA, and Taylor LS (2010a) Crystallization tendency of Eur J Pharm Sci 28:1–6.
active pharmaceutical ingredients following rapid solvent evaporation: classifica- Verheyen S, Blaton N, Kinget R, and Van den Mooter G (2002) Mechanism of in-
tion and comparison with crystallization tendency from undercooled melts. creased dissolution of diazepam and temazepam from polyethylene glycol 6000
J Pharm Sci 99:3826–3838. solid dispersions. Int J Pharm 249:45–58.
Van Eerdenbrugh B, Van den Mooter G, and Augustijns P (2008) Top-down pro- Verreck G, Chun I, Peeters J, Rosenblatt J, and Brewster ME (2003a) Preparation
duction of drug nanocrystals: nanosuspension stabilization, miniaturization and and characterization of nanofibers containing amorphous drug dispersions gener-
transformation into solid products. Int J Pharm 364:64–75. ated by electrostatic spinning. Pharm Res 20:810–817.
Van Eerdenbrugh B, Van Speybroeck M, Mols R, Houthoofd K, Martens JA, Froyen Verreck G, Six K, Van den Mooter G, Baert L, Peeters J, and Brewster ME (2003b)
L, Van Humbeeck J, Augustijns P, and Van den Mooter G (2009) Itraconazole/ Characterization of solid dispersions of itraconazole and hydroxypropylmethylcellulose
TPGS/Aerosil200 solid dispersions: characterization, physical stability and in vivo prepared by melt extrusion. Part I. Int J Pharm 251:165–174.
performance. Eur J Pharm Sci 38:270–278. Vialpando M, Aerts A, Persoons J, Martens J, and Van Den Mooter G (2011) Eval-
Van Eerdenbrugh B, Vermant J, Martens JA, Froyen L, Humbeeck JV, Van den uation of ordered mesoporous silica as a carrier for poorly soluble drugs: influence
Mooter G, and Augustijns P (2010b) Solubility increases associated with crystalline of pressure on the structure and drug release. J Pharm Sci 100:3411–3420.
drug nanoparticles: methodologies and significance. Mol Pharm 7:1858–1870. Viernstein H and Stumpf C (1992) Similar central actions of intravenous methohexitone
Van Hees T, Piel G, Evrard B, Otte X, Thunus L, and Delattre L (1999) Application of suspension and solution in the rabbit. J Pharm Pharmacol 44:66–68.
supercritical carbon dioxide for the preparation of a piroxicam-beta-cyclodextrin Vilhelmsen T, Eliasen H, and Schaefer T (2005) Effect of a melt agglomeration
inclusion compound. Pharm Res 16:1864–1870. process on agglomerates containing solid dispersions. Int J Pharm 303:132–142.
Van Speybroeck M, Barillaro V, Thi TD, Mellaerts R, Martens J, Van Humbeeck J, Villaverde J, Morillo E, Pérez-Martínez JI, Ginés JM, and Maqueda C (2004) Prep-
Vermant J, Annaert P, Van den Mooter G, and Augustijns P (2009) Ordered aration and characterization of inclusion complex of norflurazon and beta-
mesoporous silica material SBA-15: a broad-spectrum formulation platform for cyclodextrin to improve herbicide formulations. J Agric Food Chem 52:864–869.
poorly soluble drugs. J Pharm Sci 98:2648–2658. Vippagunta SR, Brittain HG, and Grant DJW (2001) Crystalline solids. Adv Drug
Van Speybroeck M, Mellaerts R, Mols R, Thi TD, Martens JA, Van Humbeeck J, Deliv Rev 48:3–26.
Annaert P, Van den Mooter G, and Augustijns P (2010a) Enhanced absorption Visalakshi NA, Mariappan TT, Bhutani H, and Singh S (2005) Behavior of moisture
of the poorly soluble drug fenofibrate by tuning its release rate from ordered gain and equilibrium moisture contents (EMC) of various drug substances and
mesoporous silica. Eur J Pharm Sci 41:623–630. correlation with compendial information on hygroscopicity and loss on drying.
van Speybroeck M, Mellaerts R, Thi TD, Martens JA, Van Humbeeck J, Annaert Pharm Dev Technol 10:489–497.
P, Van den Mooter G, and Augustijns P (2011) Preventing release in the acidic Viscomi GC, Campana M, Barbanti M, Grepioni F, Polito M, Confortini D, Rosini G,
environment of the stomach via occlusion in ordered mesoporous silica enhances Righi P, Cannata V, and Braga D (2008) Crystal forms of rifaximin and their effect
the absorption of poorly soluble weakly acidic drugs. J Pharm Sci 100: on pharmaceutical properties. CrystEngComm 10:1074–1081.
4864–4876. Vishweshwar P, McMahon JA, Bis JA, and Zaworotko MJ (2006) Pharmaceutical
Van Speybroeck M, Mols R, Mellaerts R, Thi TD, Martens JA, Van Humbeeck J, co-crystals. J Pharm Sci 95:499–516.
Annaert P, Van den Mooter G, and Augustijns P (2010b) Combined use of ordered Vogt M, Kunath K, and Dressman JB (2008) Dissolution enhancement of fenofibrate
mesoporous silica and precipitation inhibitors for improved oral absorption of the by micronization, cogrinding and spray-drying: comparison with commercial
poorly soluble weak base itraconazole. Eur J Pharm Biopharm 75:354–365. preparations. Eur J Pharm Biopharm 68:283–288.
Van Speybroeck M, Williams HD, Nguyen TH, Anby MU, Porter CJH, Volcheck GW and Van Dellen RG (1998) Anaphylaxis to intravenous cyclosporine and
and Augustijns P (2012) Incomplete desorption of liquid excipients reduces the in tolerance to oral cyclosporine: case report and review. Ann Allergy Asthma
vitro and in vivo performance of self-emulsifying drug delivery systems solidified Immunol 80:159–163.
by adsorption onto an inorganic mesoporous carrier. Mol Pharm 9:2750–2760. Völgyi G, Baka E, Box KJ, Comer JEA, and Takács-Novák K (2010) Study of pH-
van Stam J, De Feyter S, De Schryver FC, and Evans CH (1996) 2-Naphthol com- dependent solubility of organic bases: revisit of Henderson-Hasselbalch relation-
plexation by beta-cyclodextrin: influence of added short linear alcohols. J Phys ship. Anal Chim Acta 673:40–46.
Chem 100:19959–19966. Vonarbourg A, Passirani C, Saulnier P, and Benoit JP (2006) Parameters influencing
van ’t Klooster G, Hoeben E, Borghys H, Looszova A, Bouche MP, van Velsen F, the stealthiness of colloidal drug delivery systems. Biomaterials 27:4356–4373.
and Baert L (2010) Pharmacokinetics and disposition of rilpivirine (TMC278) Vyazovkin S and Dranca I (2005) Physical stability and relaxation of amorphous
nanosuspension as a long-acting injectable antiretroviral formulation. Antimicrob indomethacin. J Phys Chem B 109:18637–18644.
Agents Chemother 54:2042–2050. Walkling WD, Reynolds BE, Fegely BJ, and Janicki CA (1983) Xilobam: effect of salt
van Zuylen L, Gianni L, Verweij J, Mross K, Brouwer E, Loos WJ, and Sparreboom A form on pharmaceutical properties. Drug Dev Ind Pharm 9:809–819.
(2000) Inter-relationships of paclitaxel disposition, infusion duration and cremophor Walsh RDB, Bradner MW, Fleischman S, Morales LA, Moulton B, Rodríguez-
EL kinetics in cancer patients. Anticancer Drugs 11:331–337. Hornedo N, and Zaworotko MJ (2003) Crystal engineering of the composition of
van Zuylen L, Verweij J, and Sparreboom A (2001) Role of formulation vehicles in pharmaceutical phases. Chem Commun (Camb) 2:186–187.
taxane pharmacology. Invest New Drugs 19:125–141. Walton VC, Howlett MR, and Selzer GB (1970) Anhydrotetracycline and
Variankaval N, Wenslow R, Murry J, Hartman R, Helmy R, Kwong E, Clas SD, 4-epianhydrotetracycline in market tetracyclines and aged tetracycline products.
Dalton C, and Santos I (2006) Preparation and solid-state characterization of J Pharm Sci 59:1160–1164.
498 Williams et al.

Wandel C, Kim RB, and Stein CM (2003) “Inactive” excipients such as Cremophor White KL, Nguyen G, Charman WN, Edwards GA, Faassen WA, and Porter CJ (2009)
can affect in vivo drug disposition. Clin Pharmacol Ther 73:394–396. Lymphatic transport of methylnortestosterone undecanoate (MU) and the bio-
Wang F, Hui H, Barnes TJ, Barnett C, and Prestidge CA (2010a) Oxidized availability of methylnortestosterone are highly sensitive to the mass of coadminis-
mesoporous silicon microparticles for improved oral delivery of poorly soluble tered lipid after oral administration of MU. J Pharmacol Exp Ther 331:700–709.
drugs. Mol Pharm 7:227–236. White RD, Wong J, Kipp J, Barber T, Glosson J, Kerzee J, Goud N, Cammack JN,
Wang L, Cui FD, and Sunada H (2006) Preparation and evaluation of solid dis- and Rabinow B (2003) Pre-clinical evaluation of itraconazole nanosuspension for
persions of nitrendipine prepared with fine silica particles using the melt-mixing intravenous injection. Toxicol Sci 72:51–59.
method. Chem Pharm Bull (Tokyo) 54:37–43. Wiernik PH, Schwartz EL, Einzig A, Strauman JJ, Lipton RB, and Dutcher JP (1987)
Wang YC, Liu ZP, Zhang DR, Gao XH, Zhang XY, Duan CX, Jia LJ, Feng FF, Huang Phase I trial of taxol given as a 24-hour infusion every 21 days: responses observed
YJ, and Shen YM, et al. (2011a) Development and in vitro evaluation of deacety in metastatic melanoma. J Clin Oncol 5:1232–1239.
mycoepoxydiene nanosuspension. Colloids Surf B Biointerfaces 83:189–197. Williams AC, Cooper VB, Thomas L, Griffith L J, Petts CR, and Booth SW (2004)
Wang YC, Zhang DR, Liu ZP, Liu GP, Duan CX, Jia L J, Feng FF, Zhang XY, Shi YQ, Evaluation of drug physical form during granulation, tabletting and storage. Int J
and Zhang Q (2010b) In vitro and in vivo evaluation of silybin nanosuspensions for Pharm 275:29–39.
oral and intravenous delivery. Nanotechnology 21:155104. Williams HD, Anby MU, Sassene PJ, Kleberg K, Bakala-N’Goma J-C, Calderone M,
Wang YC and Kowal RR (1980) Review of excipients and pH’s for parenteral products Jannin V, Igonin A, Partheil A, Marchaud D, and Jule E, et al. (2012b) Toward the
used in the United States. J Parenter Drug Assoc 34:452–462. establishment of standardized in vitro tests for lipid-based formulations. Part 2. The
Wang YL, Li XM, Wang LY, Xu YL, Cheng XD, and Wei P (2011b) Formulation and effect of bile salt concentration and drug loading on the performance of type I, II,
pharmacokinetic evaluation of a paclitaxel nanosuspension for intravenous de- IIIA, IIIB, and IV formulations during in vitro digestion. Mol Pharm 9:3286–3300.
livery. Int J Nanomedicine 6:1497–1507. Williams HD, Sassene PJ, Kleberg K, Bakala-N’Goma J-C, Calderone M, Jannin V,
Wang Z, Sun J, Wang Y, Liu X, Liu Y, Fu Q, Meng P, and He Z (2010c) Solid self- Igonin A, Partheil A, Marchaud D, and Jule E, et al.. (2012) Toward the estab-
emulsifying nitrendipine pellets: preparation and in vitro/in vivo evaluation. Int J lishment of standardized in vitro tests for lipid-based formulations. Part 1. Method
Pharm 383:1–6. parameterization and comparison of in vitro digestion profiles across a range of
Wang ZR, Burrell LS, and Lambert WJ (2002) Solubility of E2050 at various pH: representative formulations. J Pharmaceutical Sci 101:3360–3380.
a case in which apparent solubility is affected by the amount of excess solid. Wilson FA, Sallee VL, and Dietschy JM (1971) Unstirred water layers in intestine:
J Pharm Sci 91:1445–1455. rate determinant of fatty acid absorption from micellar solutions. Science 174:
Wanka G, Hoffmann H, and Ulbricht W (1994) Phase-diagrams and aggregation 1031–1033.
behavior of poly(oxyethylene)-poly(oxypropylene)-poly(oxyethylene) triblock copoly- Windbergs M, Gueres S, Strachan CJ, and Kleinebudde P (2010) Two-step solid lipid
mers in aqueous-solutions. Macromolecules 27:4145–4159. extrusion as a process to modify dissolution behavior. AAPS PharmSciTech 11:2–8.
Ward GH and Yalkowsky SH (1993) Studies in phlebitis. IV. Injection rate and Windheuser JJ and Higuchi T (1963) Effect of various compounds on the rate of
amiodarone-induced phlebitis. J Parenter Sci Technol 47:40–43. thiamine hydrolysis. J Pharm Sci 52:557–559.
Ware EC and Lu DR (2004) An automated approach to salt selection for new unique Wissing SA, Kayser O, and Müller RH (2004) Solid lipid nanoparticles for parenteral
trazodone salts. Pharm Res 21:177–184. drug delivery. Adv Drug Deliv Rev 56:1257–1272.
Warnakula S, Hsieh J, Adeli K, Hussain MM, Tso P, and Proctor SD (2011) New Won DH, Kim MS, Lee S, Park JS, and Hwang SJ (2005) Improved physicochemical
insights into how the intestine can regulate lipid homeostasis and impact vascular characteristics of felodipine solid dispersion particles by supercritical anti-solvent
disease: frontiers for new pharmaceutical therapies to lower cardiovascular disease precipitation process. Int J Pharm 301:199–208.
risk. Can J Cardiol 27:183–191. Wondraczek L and Mauro JC (2009) Advancing glasses through fundamental re-
Warren DB, Anby MU, Hawley A, and Boyd BJ (2011) Real time evolution of liquid search. J Eur Ceram Soc 29:1227–1234.
crystalline nanostructure during the digestion of formulation lipids using syn- Wong J, Brugger A, Khare A, Chaubal M, Papadopoulos P, Rabinow B, Kipp J,
chrotron small-angle X-ray scattering. Langmuir 27:9528–9534. and Ning J (2008) Suspensions for intravenous (IV) injection: a review of de-
Warren DB, Benameur H, Porter CJH, and Pouton CW (2010) Using polymeric velopment, preclinical and clinical aspects. Adv Drug Deliv Rev 60:939–954.
precipitation inhibitors to improve the absorption of poorly water-soluble drugs: Woodcock DM, Jefferson S, Linsenmeyer ME, Crowther PJ, Chojnowski GM,
a mechanistic basis for utility. J Drug Target 18:704–731. Williams B, and Bertoncello I (1990) Reversal of the multidrug resistance pheno-
Wasan KM, Brocks DR, Lee SD, Sachs-Barrable K, and Thornton SJ (2008) Impact of type with cremophor EL, a common vehicle for water-insoluble vitamins and drugs.
lipoproteins on the biological activity and disposition of hydrophobic drugs: Cancer Res 50:4199–4203.
implications for drug discovery. Nat Rev Drug Discov 7:84–99. Woodcock DM, Linsenmeyer ME, Chojnowski G, Kriegler AB, Nink V, Webster LK,
Washington L, Cook GA, and Mansbach CM 2nd (2003) Inhibition of carnitine and Sawyer WH (1992) Reversal of multidrug resistance by surfactants. Br J
palmitoyltransferase in the rat small intestine reduces export of triacylglycerol Cancer 66:62–68.
into the lymph. J Lipid Res 44:1395–1403. Wu CY and Benet LZ (2005) Predicting drug disposition via application of BCS:
Washington N, Steele RJC, Jackson SJ, Bush D, Mason J, Gill DA, Pitt K, transport/absorption/elimination interplay and development of a biopharmaceutics
and Rawlins DA (2000) Determination of baseline human nasal pH and the effect of drug disposition classification system. Pharm Res 22:11–23.
intranasally administered buffers. Int J Pharm 198:139–146. Wu K, Li J, Wang WN, and Winstead DA (2009) Formation and characterization of
Watanabe T, Nakayama Y, Naito M, Oh-hara T, Itoh Y, and Tsuruo T (1996) solid dispersions of piroxicam and polyvinylpyrrolidone using spray drying and
Cremophor EL reversed multidrug resistance in vitro but not in tumor-bearing precipitation with compressed antisolvent. J Pharm Sci 98:2422–2431.
mouse models. Anticancer Drugs 7:825–832. Wu LB, Zhang J, and Watanabe W (2011) Physical and chemical stability of drug
Watanabe T, Wakiyama N, Usui F, Ikeda M, Isobe T, and Senna M (2001) Stability of nanoparticles. Adv Drug Deliv Rev 63:456–469.
amorphous indomethacin compounded with silica. Int J Pharm 226:81–91. Wu WJ and Nancollas GH (1998) A new understanding of the relationship between
Watari N, Funaki T, Aizawa K, and Kaneniwa N (1983) Nonlinear assessment of solubility and particle size. J Solution Chem 27:521–531.
nitrofurantoin bioavailability in rabbits. J Pharmacokinet Biopharm 11: Wu X, Zhou M, Huang LS, Wetterau J, and Ginsberg HN (1996) Demonstration of
529–545. a physical interaction between microsomal triglyceride transfer protein and apo-
Webster L, Linsenmeyer M, Millward M, Morton C, Bishop J, and Woodcock D (1993) lipoprotein B during the assembly of ApoB-containing lipoproteins. J Biol Chem
Measurement of cremophor EL following taxol: plasma levels sufficient to reverse 271:10277–10281.
drug exclusion mediated by the multidrug-resistant phenotype. J Natl Cancer Inst Wu Y, Loper A, Landis E, Hettrick L, Novak L, Lynn K, Chen C, Thompson K,
85:1685–1690. Higgins R, and Batra U, et al. (2004) The role of biopharmaceutics in the de-
Wei Z, Hao JG, Yuan S, Li YJ, Juan W, Sha XY, and Fang XL (2009) Paclitaxel- velopment of a clinical nanoparticle formulation of MK-0869: a Beagle dog model
loaded Pluronic P123/F127 mixed polymeric micelles: formulation, optimization predicts improved bioavailability and diminished food effect on absorption in hu-
and in vitro characterization. Int J Pharm 376:176–185. man. Int J Pharm 285:135–146.
Weiss RB, Donehower RC, Wiernik PH, Ohnuma T, Gralla RJ, Trump DL, Baker JR Wulff-Pérez M, de Vicente J, Martín-Rodríguez A, and Gálvez-Ruiz MJ (2012) Con-
Jr, Van Echo DA, Von Hoff DD, and Leyland- Jones B (1990) Hypersensitivity trolling lipolysis through steric surfactants: new insights on the controlled degra-
reactions from taxol. J Clin Oncol 8:1263–1268. dation of submicron emulsions after oral and intravenous administration. Int J
Weng Larsen S and Larsen C (2009) Critical factors influencing the in vivo perfor- Pharm 423:161–166.
mance of long-acting lipophilic solutions: impact on in vitro release method design. Wulff M, Alden M, and Craig DQM (1996) An investigation into the critical surfac-
AAPS J 11:762–770. tant concentration for solid solubility of hydrophobic drug in different polyethylene
Werle M (2008) Polymeric and low molecular mass efflux pump inhibitors for oral glycols. Int J Pharm 142:189–198.
drug delivery. J Pharm Sci 97:60–70. Xiang TX and Anderson BD (2002) Stable supersaturated aqueous solutions of
Weuts I, Kempen D, Decorte A, Verreck G, Peeters J, Brewster M, and Van den silatecan 7-t-butyldimethylsilyl-10-hydroxycamptothecin via chemical conversion in
Mooter G (2005) Physical stability of the amorphous state of loperamide and two the presence of a chemically modified beta-cyclodextrin. Pharm Res 19:1215–1222.
fragment molecules in solid dispersions with the polymers PVP-K30 and PVP- Xiong RL, Lu WG, Li J, Wang PQ, Xu R, and Chen TT (2008) Preparation and
VA64. Eur J Pharm Sci 25:313–320. characterization of intravenously injectable nimodipine nanosuspension. Int J
Weuts I, Kempen D, Six K, Peeters J, Verreck G, Brewster M, and Van den Mooter G Pharm 350:338–343.
(2003) Evaluation of different calorimetric methods to determine the glass tran- Yalkowsky SH (1999) Solubility and Solubilization in Aqueous Media, Oxford Uni-
sition temperature and molecular mobility below T(g) for amorphous drugs. Int J versity Press, New York.
Pharm 259:17–25. Yalkowsky SH, Flynn GL, and Amidon GL (1972) Solubility of nonelectrolytes in
Weuts I, Van Dycke F, Voorspoels J, De Cort S, Stokbroekx S, Leemans R, Brewster polar solvents. J Pharm Sci 61:983–984.
ME, Xu D, Segmuller B, Turner YT, and Roberts CJ, et al. (2011) Physicochemical Yalkowsky SH, Krzyzaniak JF, and Ward GH (1998) Formulation-related problems
properties of the amorphous drug, cast films, and spray dried powders to predict associated with intravenous drug delivery. J Pharm Sci 87:787–796.
formulation probability of success for solid dispersions: etravirine. J Pharm Sci Yalkowsky SH, Valvani SC, and Johnson BW (1983) In vitro method for detecting
100:260–274. precipitation of parenteral formulations after injection. J Pharm Sci 72:1014–1017.
White DG, Story MJ, and Barnwell SG (1991) An experimental animal model for Yamagata T, Kusuhara H, Morishita M, Takayama K, Benameur H, and Sugiyama Y
studying the effects of a novel lymphatic drug delivery system for propranolol. Int J (2007a) Effect of excipients on breast cancer resistance protein substrate uptake
Pharm 69:169–174. activity. J Control Release 124:1–5.
Strategies for Low Drug Solubility 499
Yamagata T, Kusuhara H, Morishita M, Takayama K, Benameur H, and Sugiyama Y Yunomae K, Arima H, Hirayama F, and Uekama K (2003) Involvement of cholesterol
(2007b) Improvement of the oral drug absorption of topotecan through the in- in the inhibitory effect of dimethyl-beta-cyclodextrin on P-glycoprotein and MRP2
hibition of intestinal xenobiotic efflux transporter, breast cancer resistance protein, function in Caco-2 cells. FEBS Lett 536:225–231.
by excipients. Drug Metab Dispos 35:1142–1148. Zakir F, Sharma H, Kaur K, Malik B, Vaidya B, Goyal AK, and Vyas SP (2011)
Yamagata T, Morishita M, Kusuhara H, Takayama K, Benameur H, and Sugiyama Y Nanocrystallization of poorly water soluble drugs for parenteral administration.
(2009) Characterization of the inhibition of breast cancer resistance protein- J Biomed Nanotechnol 7:127–129.
mediated efflux of mitoxantrone by pharmaceutical excipients. Int J Pharm 370: Zana R (1995) Aqueous surfactant-alcohol systems: a review. Adv Colloid Interface
216–219. Sci 57:1–64.
Yamamoto A, Taniguchi T, Rikyuu K, Tsuji T, Fujita T, Murakami M, and Muranishi Zannou EA, Ji Q, Joshi YM, and Serajuddin ATM (2007) Stabilization of the maleate
S (1994) Effects of various protease inhibitors on the intestinal absorption and salt of a basic drug by adjustment of microenvironmental pH in solid dosage form.
degradation of insulin in rats. Pharm Res 11:1496–1500. Int J Pharm 337:210–218.
Yamashita K, Nakate T, Okimoto K, Ohike A, Tokunaga Y, Ibuki R, Higaki K, Zaslavsky BY, Ossipov NN, and Rogozhin SV (1978) Action of surface-active sub-
and Kimura T (2003) Establishment of new preparation method for solid dispersion stances of biological membranes. III. Comparison of hemolytic activity of ionic and
formulation of tacrolimus. Int J Pharm 267:79–91. nonionic surfactants. Biochim Biophys Acta 510:151–159.
Yamashita T, Ozaki S, and Kushida I (2011) Solvent shift method for anti-precipitant Zhang GGZ, Henry RF, Borchardt TB, and Lou XC (2007) Efficient co-crystal
screening of poorly soluble drugs using biorelevant medium and dimethyl sulfox- screening using solution-mediated phase transformation. J Pharm Sci 96:990–995.
ide. Int J Pharm 419:170–174. Zhang GGZ, Law D, Schmitt EA, and Qiu YH (2004) Phase transformation consid-
Yáñez JA, Wang SW, Knemeyer IW, Wirth MA, and Alton KB (2011) Intestinal erations during process development and manufacture of solid oral dosage forms.
lymphatic transport for drug delivery. Adv Drug Deliv Rev 63:923–942. Adv Drug Deliv Rev 56:371–390.
Yang G, Jain N, and Yalkowsky SH (2004) Combined effect of SLS and (SBE)(7M)- Zhang H, Yao M, Morrison RA, and Chong S (2003) Commonly used surfactant,
beta-CD on the solubilization of NSC-639829. Int J Pharm 269:141–148. Tween 80, improves absorption of P-glycoprotein substrate, digoxin, in rats. Arch
Yang TF, Chen CN, Chen MC, Lai CH, Liang HF, and Sung HW (2007) Shell- Pharm Res 26:768–772.
crosslinked Pluronic L121 micelles as a drug delivery vehicle. Biomaterials 28: Zhang WD, Zhang C, Liu RH, Li HL, Zhang JT, Mao C, Moran S, and Chen CL (2006)
725–734. Preclinical pharmacokinetics and tissue distribution of a natural cardioprotective
Yang Y, Wang JC, Zhang X, Lu WL, and Zhang Q (2009) A novel mixed micelle gel agent astragaloside IV in rats and dogs. Life Sci 79:808–815.
with thermo-sensitive property for the local delivery of docetaxel. J Control Release Zhang Z and Feng SS (2006) Self-assembled nanoparticles of poly(lactide)-vitamin E
135:175–182. TPGS copolymers for oral chemotherapy. Int J Pharm 324:191–198.
Yao WW, Bai TC, Sun JP, Zhu CW, Hu J, and Zhang HL (2005) Thermodynamic Zhao YY, Inbar P, Chokshi HP, Malick AW, and Choi DS (2011) Prediction of the
properties for the system of silybin and poly(ethylene glycol) 6000. Thermochim thermal phase diagram of amorphous solid dispersions by Flory-Huggins theory.
Acta 437:17–20. J Pharm Sci 100:3196–3207.
Yeh PY, Smith PL, and Ellens H (1994) Effect of medium-chain glycerides on physi- Zheng W, Jain A, Papoutsakis D, Dannenfelser RM, Panicucci R, and Garad S (2012)
ological properties of rabbit intestinal epithelium in vitro. Pharm Res 11:1148–1154. Selection of oral bioavailability enhancing formulations during drug discovery.
Yi T, Wan J, Xu H, and Yang X (2008) A new solid self-microemulsifying formulation Drug Dev Ind Pharm 38:235–247.
prepared by spray-drying to improve the oral bioavailability of poorly water soluble Zhou D, Zhang GGZ, Law D, Grant DJW, and Schmitt EA (2008) Thermodynamics,
drugs. Eur J Pharm Biopharm 70:439–444. molecular mobility and crystallization kinetics of amorphous griseofulvin. Mol
Yokoi Y, Yonemochi E, and Terada K (2005) Effects of sugar ester and hydroxypropyl Pharm 5:927–936.
methylcellulose on the physicochemical stability of amorphous cefditoren pivoxil in Zhou DL, Zhang GGZ, Law D, Grant DJW, and Schmitt EA (2002) Physical stability
aqueous suspension. Int J Pharm 290:91–99. of amorphous pharmaceuticals: importance of configurational thermodynamic
Yoo JS, Ning SM, Pantuck CB, Pantuck EJ, and Yang CS (1991) Regulation of he- quantities and molecular mobility. J Pharm Sci 91:1863–1872.
patic microsomal cytochrome P450IIE1 level by dietary lipids and carbohydrates in Zhou HH, Goldman M, Wu J, Woestenborghs R, Hassell AE, Lee P, Baruch A, Pesco-
rats. J Nutr 121:959–965. Koplowitz L, Borum J, and Wheat L J (1998) A pharmacokinetic study of in-
Yoo SD, Lee SH, Kang EH, Jun H, Jung JY, Park JW, and Lee KH (2000) Bio- travenous itraconazole followed by oral administration of itraconazole capsules in
availability of itraconazole in rats and rabbits after administration of tablets patients with advanced human immunodeficiency virus infection. J Clin Pharmacol
containing solid dispersion particles. Drug Dev Ind Pharm 26:27–34. 38:593–602.
Yoshioka M, Hancock BC, and Zografi G (1994) Crystallization of indomethacin from Zhu HJ, Meserve K, and Floyd A (2002) Preformulation studies for an ultrashort-
the amorphous state below and above its glass transition temperature. J Pharm acting neuromuscular blocking agent GW280430A. I. Buffer and cosolvent effects
Sci 83:1700–1705. on the solution stability. Drug Dev Ind Pharm 28:135–142.
Yoshioka M, Hancock BC, and Zografi G (1995) Inhibition of indomethacin crystal- Zia V, Rajewski RA, and Stella VJ (2001) Effect of cyclodextrin charge on complex-
lization in poly(vinylpyrrolidone) coprecipitates. J Pharm Sci 84:983–986. ation of neutral and charged substrates: comparison of (SBE)7M-beta-CD to
Yoshioka S, Aso Y, and Kojima S (1999) The effect of excipients on the HP-beta-CD. Pharm Res 18:667–673.
molecular mobility of lyophilized formulations, as measured by glass transition Ziller KH and Rupprecht H (1988) Control of crystal-growth in drug suspensions. 1.
temperature and NMR relaxation-based critical mobility temperature. Pharm Res Design of a control unit and application to acetaminophen suspensions. Drug Dev
16:135–140. Ind Pharm 14:2341–2370.
Young TJ, Mawson S, Johnston KP, Henriksen IB, Pace GW, and Mishra AK (2000) Zimmermann A, Tian F, Lopez de Diego H, Ringkjøbing Elema M, Rantanen J,
Rapid expansion from supercritical to aqueous solution to produce submicron Müllertz A, and Hovgaard L (2009) Influence of the solid form of siramesine hy-
suspensions of water-insoluble drugs. Biotechnol Prog 16:402–407. drochloride on its behavior in aqueous environments. Pharm Res 26:846–854.
Yu H, Hu YQ, Ip FCF, Zuo Z, Han YF, and Ip NY (2011a) Intestinal transport of Zimmermann E, Müller RH, and Mäder K (2000) Influence of different parameters on
bis(12)-hupyridone in Caco-2 cells and its improved permeability by the surfactant reconstitution of lyophilized SLN. Int J Pharm 196:211–213.
Brij-35. Biopharm Drug Dispos 32:140–150. Zordan-Nudo T, Ling V, Liu Z, and Georges E (1993) Effects of nonionic detergents on
Yu HZ, Han SF, Li P, Zhu CL, Zhang XX, Gan L, and Gan Y (2011b) An examination P-glycoprotein drug binding and reversal of multidrug resistance. Cancer Res 53:
of the potential effect of lipids on the first-pass metabolism of the lipophilic drug 5994–6000.
anethol trithione. J Pharm Sci 100:5048–5058. Zuidema J, Kadir F, Titulaer HAC, and Oussoren C (1994) Release and absorption
Yu L (2001) Amorphous pharmaceutical solids: preparation, characterization and rates of intramuscularly and subcutaneously injected pharmaceuticals (II). Int J
stabilization. Adv Drug Deliv Rev 48:27–42. Pharm 105:189–207.
Yu LX, Amidon GL, Polli JE, Zhao H, Mehta MU, Conner DP, Shah VP, Lesko LJ, zur Mühlen A, Schwarz C, and Mehnert W (1998) Solid lipid nanoparticles (SLN) for
Chen ML, and Lee VHL, et al. (2002) Biopharmaceutics classification system: the controlled drug delivery-drug release and release mechanism. Eur J Pharm Bio-
scientific basis for biowaiver extensions. Pharm Res 19:921–925. pharm 45:149–155.

S-ar putea să vă placă și