Sunteți pe pagina 1din 5

Research: Science and Education

Understanding, Deriving, and Computing Buffer Capacity


Edward T. Urbansky* and Michael R. Schock**
National Risk Management Research Laboratory, Water Supply and Water Resources Division, U.S. Environmental
Protection Agency, Cincinnati, OH 45268-0001; *Urbansky.Edward@EPA.gov; **Schock.Michael@EPA.gov

We dedicate this paper to the memory of aquatic chemist beyond the scope of this article, the activity coefficient cannot
Werner Stumm and to all those who helped to lay the be determined. In practice, pH is defined operationally in
foundations of aqueous equilibrium chemistry. terms of a Nernstian relationship (eq 2) between an electrodes
response (measured potential) in a solution of unknown pH
In the past two years more than 600 papers have been
and its response in a solution of known pH. In both cases,
published involving buffer capacity. It appears in varied
the potential is measured relative to a standard reference half-
disciplines, including analytical, environmental, geo-, and
cell, for example Hg/Hg2Cl2 or Ag/AgCl electrodes:
biochemistry, physiology, medicine, dentistry, and agriculture.
The following indicates the range of topics in the recent F E unk E std
literature: acid rain (1), pollution control (2), metal interactions pHunk = pHstd + (2)
in fish (3), cellular metabolism (4), drug solubility (5), muscle ln 10 RT
function (67), thyroid function (8), tooth decay (911), plant
nutrient uptake (12), soil lime application (13), potentiometry where pHunk is the calculated pH of the unknown solution;
(14), and corrosion in potable water systems (1516 ). From pHstd is the defined pH of the standard solution; Eunk is the
these examples, it is clear that buffer capacity finds application measured potential of the unknown solution relative to the
throughout the sciences. reference electrode; Estd is the measured potential of the
Despite this importance, a well-known environmental standard solution; R is the universal gas constant (8.314 J
chemistry text omits the topic (17), and a volume on corrosion mol!1 K!1); and F is the Faraday constant, 96,485 C (mol e)!1.
has limited coverage (18). Although mass balance and charge In the United States, standard buffers of known pH are
balance are covered in quantitative analysis textbooks (19 established by the National Institute of Standards and
21), buffer capacity is treated minimally. Texts that treat equi- Technology (NIST), and NIST-traceable buffers are routinely
librium and buffer capacity more rigorously are often reserved used in American laboratories.
for upper-level or specialty chemistry courses (2229). It is also possible to directly measure the proton molar
Donald Van Slykes classic authoritative work (30) was concentration, represented as [H+]. This is done by the cali-
published in 1922, several years after that of better known brating electrodes through titrations of strong acid and strong
Henderson and Hasselbalch. Van Slykes specialty was blood base (e.g., HClO4 and NaOH) at constant ionic strength1 in
chemistry, where carbon dioxide equilibria dominate the acid a strong electrolyte (depending on the application, usually
base chemistry. Since that time, attempts have been made to KNO3, NaCl, LiClO4 or NaClO4) and while sparging with
clarify the definition of buffer capacity (31) or provide an a minimally soluble unreactive gas (e.g., Ar) to keep out CO2.
approach for its computation (32). Others have expanded Van To prevent evaporative loss, the gas is presaturated with water
Slykes work, applying the concept of buffer capacity to various vapor. Proton concentrations are often used in kinetic studies
areas (27, 29, 3336). Fortunately, the advent of spreadsheets where it is desirable to accurately quantitate all species in
has made short work of the lengthy, involved calculations re- molar concentrations. Equilibrium constants are expressed in
quired for effective use of buffer capacity (1, 3739), and molar concentrations rather than activities, but they are valid
JCE Software itself has produced a software package that only for a specific ionic strength and medium, for example,
includes buffer capacity calculation (40). 0.50 M NaClO4. Other than seawater, natural and potable
waters tend to have low ionic strength; thus, this approach is
limited in applicability even though it is always valid. Activity-
Definition of pH based equilibrium constants are generally measured using an
Understanding pH is a prerequisite for understanding inert strong electrolyte to control ionic strength. After deter-
buffer capacity. The term pH is formally defined by eq 1: mination of the molar-concentration-based values, a plot of
K against is used to extrapolate to zero ionic strength (infinite
H+ m H+ dilution).
pH = !log H+ = !log (1) In common usage, a pH meter estimates {H+}. As solu-
m H+
tions become more dilute and ionic strength approaches zero,
where {H +} represents the activity of hydrogen ion, H+ the activity coefficients for all ions approach unity; that is,
represents the activity coefficient of the hydrogen ion, m H+ lim
represents the molality of the hydrogen ion in the solution =1
0
being measured, and mH+ represents the standard state molal-
ity of the hydrogen ion, normally defined as unity. For reasons Unit activity coefficients are routinely assumed since 1
for < 0.005 M. Nevertheless, this remains an assumption;

This paper is the work product of United States government a particular situation may demand ionic strength corrections.
employees engaged in their official duties and is therefore in the By assuming unit activity coefficients, we can interchange
public domain and exempt from copyright. molalities, molarities, and activities and dispense with activity

1640 Journal of Chemical Education Vol. 77 No. 12 December 2000 JChemEd.chem.wisc.edu


Research: Science and Education

coefficients in subsequent equations. However, we must not charge balance: [H+] + [Na+] = [OH!] + [X !] (9)
forget that activity coefficients are always implicitly present. +
It is possible to make the substitution [Na ] = Cb, the con-
centration of base that must have been added. In addition,
Definition of Buffer Capacity 2 we substitute [OH!] = Kw/[H+], and solve for Cb.
For an aqueous solution, the buffer capacity is defined
Cb = Kw /[H+] [H+] + [X!] (10)
in terms of the concentration of acid or base that must be
added to influence pH. The formal definition of the buffer
To express eq 10 in terms of [X]T, we must use the mass
capacity, !, is given by eq 3:
balance (eq 9) and the acid dissociation constant, Ka =
! = dCb/d(pH) = !dCa/d(pH) (3) [H+][X!]/[HX]. The ratio of total concentration of all forms
of X to that of X!, which also represents the reciprocal of the
where Cb represents the concentration (normality) of added
fraction of X!, is given by eq 11. Subsequent substitution into
base and Ca represents the concentration of added acid. Van
eq 10 gives eq 12, which simplifies to eq 13.
Slyke expressed the definition in this form because he was
most interested in deducing [CO2] or [HCO3!] for a (pH)
X T H+
that he could measure. We retain this form for historical =1+ (11)
reasons and because it simplifies the mathematics. X
! Ka
Buffer capacity is always positive; every solution resists
pH change according to Le Chteliers principle. Because pH Kw X
goes down upon the addition of acid, a minus sign is required Cb = +
H+ + T
+ (12)
H H
when considering acid. Buffer capacity is a continuous and 1+
nonlinear function; therefore, it is a derivative. For a given Ka
value of pH, it makes no difference whether we think in terms
of a differential change in pH resulting from an infinitesimal Kw Ka
addition of acid versus an infinitesimal addition of base (other Cb = +
H+ + X T (13)
than the sign of the differential). H K a + H+
The buffer capacity can be resolved into a series of terms,
with one term for each active component. At this point, we differentiate with respect to [H+]:

! = !OH ! + !H+ + !i(all weak acids and bases) (4) dC b Kw !K a 1


+
=! 1+ X T (14)
For convenience, the first two terms (!OH ! + ! ) are often H+ dH H
+ 2
K a + H+
2
combined and referred to as the buffer capacity of water, !w.
However, !OH ! and !H+ increase upon the respective addi- Since d[H+]/d(pH) = !ln 10 [H+], we apply the chain rule
tion of strong Arrhenius bases and acids. !w should not be for composite functions to obtain the final result:
associated with Kw and does not stem only from the auto-
ionization of water.
dC b Kw + K a H+
An n-protic acid and its conjugate bases make a contri- != = ln 10 + H + X T (15)
+
bution to the overall buffer capacity that is expressible as the d pH H K a + H+
2

following equivalent statements (eqs 57):


i=n1
Equation 15 shows three distinct components of !. The hy-
!H X = X T d n 1 f H X n! (5)
droxide and hydrogen terms can be expressed as !w, and the
n d pH i=1 i weak electrolyte term as !buffer. Figure 1 shows the buffer
capacity plot for acetic acid. The logic applies likewise to a
i=n1 weak base, for example, NH3. In this case the mass balance
d n 1 i H+
i is given by [B]T = [B] + [HB+], and Ca (instead of Cb) is used
!H X = X
n T
d pH
f X n! n +
i=1
(6)
to write the charge balance: [H+] + [HB+] = Ca + [OH!].
Nonetheless, the final result obtained for ! is the same.
i=n1
+ d + i
!H X = !ln 10 H
n
X T
d H+
f X n! n +
i=1
n 1 i H (7) Diprotic Weak Acid and Its Conjugate Bases
The buffering of natural and potable waters and
where the ith cumulative protonation constant i = [HiAi!q]/
biophysiological solutions is influenced by dissolved carbon
i=n
([H+]i[Aq!]), and the fraction of Xn!, fXn! = 1/(1 + i=1 i [H+]i ).
dioxide, carbonic acid,3 hydrogen carbonate, and carbonate.
Application of these equations follows.
As a result, the buffer capacity expression for the diprotic acid,
CO2, has been derived elsewhere. Only the final results will
Monoprotic Weak Acid and Its Conjugate Base be stated here.
First consider the case of a monoprotic weak acid, such + + 2
as acetic acid and its conjugate base (e.g., the sodium salt). + 1 + 4 2 H + 1 2 H
!CO = ln 10 H CO2 T (16)
2 2 2
mass balance: [X]T = [HX] + [X ! ] (8) 1 + 1 H+ + 2 H+

JChemEd.chem.wisc.edu Vol. 77 No. 12 December 2000 Journal of Chemical Education 1641


Research: Science and Education

2
K a K a + 4K a H+ + H+
!CO = ln 10 K a H+ CO2 T 1 2 2
(17)
2 1 2 2
K a K a + K a H+ + H+
1 2 1

We note that the same equations apply to all diprotic acids


and bases, even zwitterions (e.g., glycine), for which the
initial charge balance equations are different. We leave these
derivations as exercises.

Triprotic Weak Acid and Its Conjugate Bases


This derivation is more complicated and some level of
detail is instructive.
mass balance: [X]T = [X3!] + [HX2!] + [H2X!] + [H3X] (18) Figure 1. Buffer capacity (left ordinate) for the acetic acid/acetate +
water system is calculated using eq 15. [HOAc]T = [HOAc] +
charge balance: Cb = Kw/[H+] [H+] + 3[X3!] + 2[HX2!] + [H2X!] (19) [OAc!] = 0.10 M and pKa = 4.50. A relative maximum occurs at
Limiting ourselves to the contribution of the weak electrolyte pH = pKa = 4.50. If we approximate ! using (pH) = 0.14 in eq
26, an error is introduced. The relative error, (!approx !)/!, is
system alone (i.e., excluding !w), we use the equilibrium
also shown (right ordinate). !w is not shown explicitly.
expressions and execute similar mathematical operations:
dCb(H3X)/d[H+] = [X]T S/D (20)
where S = (1 + 1[H+] + 2[H+]2 + 3[H+]3)(21 + 22[H+])
(3 + 2 1 [H+ ] + 2[H +] 2 )( 1 + 2 2[H + ] + 3 3[H + ] 2 ) and
D = (1 + 1[H+] + 2[H+]2 + 3[H+]3)2. Multiplication of
eq 20 by (! ln 10)[H+] produces the following result for the
buffer capacity contribution of this weak electrolyte system:
!H X = (ln 10)[H+][X]T N/D
3
(21)
where N = 1 + 42[H+] + ( 1 2 + 93)[H+]2 + 413[H+]3 +
23[H+]4. The function is plotted for phosphoric acid in
Figure 2.

Hydrolysis Reactions Figure 2. Buffer capacity for the phosphate system is calculated
using eq 21. [PO4]T = 0.10 M, pK a 1 = 2.15, pK a 2 = 7.20, pK a 3 =
Many metal cations, for example Al(III), are Lewis acids, 12.15; log 1 = 12.15, log 2 = 19.35, and log 3 = 21.50.
reacting with water to produce hydrogen ions. They can Contributions of H+ and OH! are shown as !w. Contribution of the
precipitate as insoluble hydroxides and oxides; accordingly, phosphate system is !(H3PO4). ! = !(H3PO4) + !w. Local maxima
it is necessary to ensure that solubility products have not been occur where the pH reaches a pKa.
exceeded by the ion products. In this case, cumulative hydrolysis
constants, rather than protonation constants are used.
plicated cases may be approximated with reasonable accuracy.
mass balance: [AlIII]T = [Al3+] + [AlOH2+] + Equilibrium constants generally have at least 15% uncer-
(22)
[Al(OH)2+] + [Al(OH)3] + [Al(OH)4!] tainty. Accordingly, even the analytic solutions are limited
charge balance: [Al(OH)4!] + [OH! ] = by experimental error. We recall the definition of buffer
(23) capacity as a derivative:
C b + [H+] + [Al(OH)2+] + 2[AlOH2+] + 3[Al3+]
We leave it to readers and their advanced students to show that dC b lim C b
!= = (25)
!Al(III) = (ln 10) [AlIII]T [H+]N/D (24) d pH pH0 pH

where N = h [H+]6 + 4h [H+]5 + (9h + h h )[H+]4 + (16h +


1 2 3 1 2 4
As a continuous function, ! can be subjected to linear
4h h )[H+]3 + (9h h + h h )[H+]2 + 4h h [H+] + h h
1 3 1 4 2 3 2 4 3 4
approximation. The exactness obtained depends on how
and D = ([H+]4 + h [H+]3 + h [H+]2 + h [H+] + h )2.
1 2 3 4
closely (pH) approaches d(pH). For 0.01 (pH) 0.1,
the speciation changes are often (but not always) small
Computation and Numerical Approximation enough. Values chosen for (pH) are best limited to 0.001;
otherwise, too many computations may be required. Buffer
Computing the buffer capacity equations is easily accom- capacity is always positive. However, choice of Ca or Cb is
plished with any commercial spreadsheet, and a text (20) arbitrary, and C is evaluated over the entire pH domain. Thus,
exists on spreadsheet use. The analytic solutions have been negative numbers correspond to changing from adding acid
the primary focus of this discussion; however, extremely com- to adding base or vice versa. For this reason, the absolute value

1642 Journal of Chemical Education Vol. 77 No. 12 December 2000 JChemEd.chem.wisc.edu


Research: Science and Education

is required for the approximation: can be used to find regions of maximal buffering when a
mixture of species is present and to estimate the overall
C b pH2 C b pH1 amount of acid or base that may be added without changing
!approx = (26)
pH2 pH1 solution pH. Such a need is common in aqueous studies and
can be encountered in many disciplines.
Note that (pH) = 1.0, the value used in Chiriac and Baleas
interpretation of the IUPAC definition (31), poorly approxi- Notes
mates the analytic solution. Figure 1 includes a plot that
illustrates the relative error in the acetic acid buffer capacity 1. Ionic strength is defined as = 12 c iz i2 for all i, where ci
curve when eq 26 is used for (pH) = 0.14. The relative error represents the concentration of an ion i and zi represents its charge
in the approximation reaches a maximum of about 20%, but (+1, !2, etc.).
the approximation itself can fall below the analytic solution 2. The terms buffer capacity, buffer intensity, and buffer index
by 10% as shown. The error can be reduced by choosing a are often used interchangeably. In general, analytical chemists have
smaller value for (pH), but at the cost of increasing the favored the first, geochemists the second, and engineers the third.
number of computations. In this paper, we will use the term buffer capacity. We do not make
the distinction made by Chiriac and Balea between buffer capacity
Buffer Capacity Extrema and buffer index (14). We feel this distinction is the result of the
ambiguous semantics of the IUPAC definition. IUPAC did not
Buffer capacity functions usually reach relative maxima specify a derivative explicitly, but rather described a change in mo-
at pH = pKa for an individual deprotonation reaction. How- lar concentration of strong acid or base relative to 1 pH unit. We
ever, when pKa < 3 or >11, the maxima begin to converge and interpret this usage is to mean a change in molarity per pH unit
a single relative maximum is eventually formed. One relative and thereby to represent the derivative conventionally thought of
minimum usually occurs at the intersection of the contribu- as buffer capacity. Certainly, Chiriac and Balea are correct in as-
tion of the most protonated acid form and the hydrogen ion serting that the normality of added base or acid to cause a pH shift
contribution. Another relative minimum usually occurs at the of 1 is not equal to the derivative, because the buffer capacity func-
intersection of the contribution of the most deprotonated alka- tion is nonlinear over such a large interval of the pH domain.
line form and the hydroxide ion contribution. One of these 3. Although carbonic acid is a real species, it is often excluded
is usually, but not always, the absolute minimum. However, from expressions because it behaves identically to carbon dioxide
it is possible for a species to have pKa values suitably far away in terms of overall acidity. Reference 19 has an excellent discussion
from 7 that the minimum between two forms coincides with of this matter.
the minimum of the water contribution. The function always
has an absolute minimum, but there are no absolute maxima. Literature Cited
The buffer capacity function is strictly decreasing for pH < 0,
since d! H+ /d(pH) = !2(ln 10)[H+]. It is strictly increasing 1. Walse, C.; Schopp, W.; Warfvinge, P. Environ. Pollut. 1997
for pH > 14, since d!OH! /d(pH) = 2(ln 10) K w /[H+]. As a 1998, 98, 253267.
practical limitation, buffer capacity is limited to solubilities, 2. Forstner, U.; Haase, I. J. Geochem. Explor. 1998, 62, 2936.
for example 50% w/w NaOH (ca. 19 M OH!) or 98% w/w 3. Playle, R. C. Sci. Total Environ. 1998, 219, 147163.
H2SO4 (ca. 19 M H+). However, neither of these represents 4. Walter, G.; Vandenborne, K.; Elliott, M.; Leigh, J. S. J. Physiol.
a typical aqueous solution and the activity of the water is (Cambridge) 1999, 519, 901910.
not unity. 5. Alvarez-Nunez, F. A.; Yalkowsky S. H. Int. J. Pharm. 1999,
The explicit analytic functions for buffer capacity are 185, 4549.
continuous functions and are continuously differentiable; 6. Juel, C. Acta Physiol. Scand. 1998, 162, 359366.
therefore, the first derivative test may be used to locate critical 7. Kemp, G. J. Magma 1997, 5, 231241.
points and thus possible relative minima and maxima. For 8. Doohan, M. M.; Gray, D. F.; Hool, L. C.; Robinson, B. G.;
example, a monoprotic acid and its conjugate base have ! Rasmussen, H. H. Am. J. Physiol. 1997, 272, H1589H1597.
given by 9. LeBell, Y.; Soderling, E.; Karjalainen, S. Scand. J. Dent. Res.
1991, 99, 505509.
! = 2(ln 10)K a[X]T [H+]([H+] Ka)/([H+] + K a) (27) 10. Larsen, M. J.; Jensen, A. F.; Madsen, D. M.; Pearce, E. I. Arch.
Because ! changes sign about pK a, a relative extremum must Oral Biol. 1999, 44, 111117.
occur at the critical point (pKa, !); in this case, it is a relative 11. Fure, S.; Lingstron, P.; Birkhed, D. J. Dent. Res. 1998, 77,
maximum. 16301637.
12. Probert, M. E.; Moody, P. W. Aust. J. Soil Res. 1998, 36,
Applications 389393.
13. Aitken, R. L.; Moody, P. W.; Dickson, T. Aust. J. Agric. Res.
Suppose the concentration of the titrant is much larger 1998, 49, 627637.
than the concentration of the species being titrated, that is, 14. Kwun, I.-S.; Kim, M. J. Food Sci. Nutr. 1993, 3, 3642.
Ctitrant >> Ctitrated . The slope of the titration curve, d(pH)/ 15. Clement, J. A.; Schock, M. R. Buffer Intensity: What It Is and
dV, is ca. 1/!. Buffer capacity is a guide to titration. This Why Its Critical for Controlling Distribution System Water
occurs, for example, if 1.0 mL of 5.0 mM CH3CO2H(aq) is Quality; Proc. Am. Water Works Assn. Water Qual. Technol. Conf.
titrated with 2.0 M NaOH(aq); such a method is used in the Nov 14, 1998, San Diego, CA (on CD-ROM).
determination of stability constants. Plots of buffer capacity 16. Antoun, E. N.; Hiltebrand, D. J.; Gruber, A. D. pH Instabil-

JChemEd.chem.wisc.edu Vol. 77 No. 12 December 2000 Journal of Chemical Education 1643


Research: Science and Education

ity in Relation to Implementation of Distribution System 26. Perrin, D. D. In Treatise on Analytical Chemistry, Vol. 2;
Corrosion Control; Proc. Am. Water Works Assn. Water Qual. Kolthoff, I. M.; Elving, P. J., Eds.; Wiley Interscience: New
Technol. Conf. Jun 1419, 1997, Atlanta, GA (on CD-ROM). York, 1979; Part 1, Section D, Chapter 21.
17. Manahan, S. E. Environmental Chemistry, 6th ed.; Lewis: New 27. Langmuir, D. Aqueous Environmental Geochemistry; Prentice-
York, 1994. Hall: Upper Saddle River, NJ, 1997.
18. Internal Corrosion of Water Distribution Systems, 2nd ed.; Coop- 28. Butler, J. N. Ionic Equilibrium: Solubility and pH Calculations;
erative Research Resport; American Water Works Association Wiley: New York, 1998.
Research Foundation: Denver, CO, 1996. 29. Stumm, W.; Morgan, J. J. Aquatic Chemistry: Chemical Equi-
19. Harris, D. Quantitative Chemical Analysis, 5th ed.; Freeman: libria and Rates in Natural Waters. Wiley: New York, 1995.
New York, 1999. 30. Van Slyke, D. D. J. Biol. Chem. 1922, 52, 525570.
20. Christian, G. D. Analytical Chemistry, 5th ed.; Wiley: New 31. Chiriac, V.; Balea, G. J. Chem. Educ. 1997, 74, 937939.
York, 1994. 32. King, D. W.; Kester, D. R. J. Chem. Educ. 1990, 67, 932933.
21. Skoog, D. A.; West, D. M.; Holler, F. J. Fundamentals of 33. Weber, W. J. Jr.; Stumm, W. J. Am. Water Works Assoc. 1963,
Analytical Chemistry, 7th ed.; Saunders: Fort Worth, TX, 1996. 55, 15531578.
22. Schwarzenbach, G. Complexometric Titrations; Irving, H. N. 34. Whitfield, M. Limnol. Oceanogr. 1974, 19, 235248.
M. H., Translator; Methuen: London, 1957. 35. Gibs, J.; Schoenberger, R. J.; Suffet, I. H. Water Res. 1982,
23. Schwarzenbach, G.; Flaschka, H. Complexometric Titration; 16, 699705.
Irving, H. N. M. H., Translator; Methuen: London, 1969. 36. Kleijn, H. F. W. Int. J. Water Pollut. 1965, 9, 401413.
24. Ringbom, A. Complexation in Analytical Chemistry: A Guide 37. Freiser, H. Concepts & Calculations in Analytical Chemistry: A
for the Critical Selection of Analytical Methods Based on Com- Spreadsheet Approach; CRC Press: Boca Raton, FL, 1992.
plexation Reactions; Wiley Interscience: New York, 1963. 38. Trussell, R. R. J. Am. Water Works Assoc. 1998, 90, 7081.
25. Ringbom, A.; Wnninen, E. In Treatise on Analytical Chemistry, 39. Okamoto, H.; Mori, K.; Ohtsuka, K.; Ohuchi, H.; Ishii, H.
Vol. 2; Kolthoff, I. M.; Elving, P. J., Eds.; Wiley Interscience: Pharm. Res. 1997, 14, 299302.
New York, 1979; Part 1, Section D, Chapter 20. 40. Ramette, R. W. Buffers Plus; JCE Software 1999, 9803W.

1644 Journal of Chemical Education Vol. 77 No. 12 December 2000 JChemEd.chem.wisc.edu

S-ar putea să vă placă și