Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Interpretation of the Ultraviolet Spectra of Natural Products: International Series of Monographs on Organic Chemistry
Interpretation of the Ultraviolet Spectra of Natural Products: International Series of Monographs on Organic Chemistry
Interpretation of the Ultraviolet Spectra of Natural Products: International Series of Monographs on Organic Chemistry
Ebook753 pages4 hours

Interpretation of the Ultraviolet Spectra of Natural Products: International Series of Monographs on Organic Chemistry

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Interpretation of the Ultraviolet Spectra of Natural Products focuses on the ultraviolet spectrum of chromophores. The book first discusses single chromophores, including absorption due to electron lone pairs in saturated systems and absorption of olefins, alkynes, carbonyl compounds, and thiocarbonyl compounds. The text also takes a look at conjugated chromosomes, such as polyenes, enynes, and conjugated azomethines. The selection also evaluates C-aromatic compounds. Topics include benzenoid and hydrocarbons; phenols and their ethers; styrenes and stilbenes; aromatic carbonyl compounds; and nitro compounds. The text also discusses O- and S- heteroaromatic compounds and N-heteroaromatic compounds. The book highlights the applications of spectrophotometry to the analysis of natural products. Topics include formation of derivatives having absorbing chromophores; reactions leading to changes in absorption of added reagents; and analyses involving transformation to products suitable for spectrophotometry. The text is a good reference for readers wanting to explore chromophores.
LanguageEnglish
Release dateOct 22, 2013
ISBN9781483137209
Interpretation of the Ultraviolet Spectra of Natural Products: International Series of Monographs on Organic Chemistry

Related to Interpretation of the Ultraviolet Spectra of Natural Products

Related ebooks

Chemistry For You

View More

Related articles

Reviews for Interpretation of the Ultraviolet Spectra of Natural Products

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Interpretation of the Ultraviolet Spectra of Natural Products - A. I. Scott

    Armstrong

    PREFACE

    THE origin of this book is a series of encounters with natural product chemists in search of their chromophores. I have tried to collect within these covers sufficient information to aid the organic, analytical and bio-chemist in his constant search for analogous or (hopefully) exact structural examples to match the ultraviolet spectrum of his current series of compounds, of either natural or synthetic origin. The template of natural product chemistry was chosen for the simple reason that most empirical work of correlative nature has stemmed from the study of materials isolated from biological sources. The emphasis is squarely laid on available correlative data, model systems and, wherever possible, working rules to aid the organic chemist in a qualitative sense. Thus, rigorous theoretical approaches to the calculation of spectra are not treated and many interesting topics which include mechanistic photochemistry, optical rotatory power, ligand field theory, fluorescence and phosphorescence are not discussed. I have simply tried to guide both novice and experienced workers to the source of the recorded spectrum of a variety of naturally occurring structures and to indicate the methods applicable to the solution of structural problems within the limits of the electron absorption spectrum.

    The first six chapters contain, largely in tabular form, the classes of chromophore of frequent recurrence in the realm of natural products. Chapter seven, kindly contributed by Dr. C.J.W.Brooks (Glasgow University) is intended to serve as a reference source for biochemical and analytical workers. The concluding chapters provide selective but, I hope, representative examples of the power of the interpretative method as outlined in the earlier part of the book. Many hu ndreds of excellent examples are available and I should welcome comments and suggestions from readers for future expansion of the second part of the work.

    It is indeed fortunate that, unlike some of the more recent spectral methods, there has been an element of timelessness in ultraviolet spectral studies. For this reason, I hope that readers and critics alike will forgive the delay in publication – a result of printing difficulties (the manuscript was completed in Spring 1961).

    Many colleagues helped in all directions and I am happy to acknowledge the encouragement of my Editor, Professor D.H.R.Barton F.R.S. whose fluent use of spectroscopic assignments gave me the idea of writing this book. At all times Professor R.A.Raphael F.R.S. offered helpful comment while Drs. J.C.D. Brand, C.J.W.Brooks, K.H.Overton and M.C. Whiting drew my attention to many important references.

    My debt to Professor W.D.Ollis is profound, since he not only read the entire manuscript as part of his summer vacation, but provided keen criticism at every step of the way. Miss P.A.Dodson prepared the index and the onerous task of typing was performed uncomplainingly by Miss B.Po B.Po. Porter.

    Authors, editors and publishers (where noted) co-operated fully in giving permission to reproduce diagrams. Finally, a word of gratitude to my wife for her patient help.

    A.I. SCOTT

    INTRODUCTION

    THE investigation of the structure and chemistry of naturally occurring compounds has intrigued the organic chemist since the initiation of his subject. The exponential increase in the rate of successful structural elucidations in the last decade owes a great deal to the empirical correlations developed for assignment of characteristic absorption in the infrared, ultraviolet and, more recently, n.m.r. spectra of organic molecules. The transparency of many groupings (and often large segments), of complex molecules to that part of the electromagnetic spectrum between 190 and 400 mμ, defined herein as the ultraviolet region, imposes per se a limitation of usefulness on the results of interpretation of absorption bands in this region. However, when taken in conjunction with the wealth of detail provided by infrared and n.m.r. bands, such a combination of data from three instrumentally accessible regions, when correctly interpreted, in many cases may lead to structural proposals of value to the natural product chemist. Such formulations may then be tested by appropriate chemical transformation with spectroscopic control. In this way, the process of degradation and structural assignment is aided by spectroscopic assignments, which admittedly are of an empirical nature at this time, but none the less are useful to the organic chemist.

    This short introduction to the interpretation of the ultraviolet spectra of some natural products is written from such an empirical standpoint in the hope that these correlations, which constitute an ever-increasing part of the language of the organic chemist, may form not only a useful working guide to the application of ultraviolet data in structural work, but may also stimulate the evolution of further assignments as they arise in this continually expanding subject of structure–spectra correlation.

    DEFINITIONS AND CONCEPTS

    Wavelength

    Electromagnetic radiation is characterized by either its wavelength (λ) or by its frequency (ν). These quantities are related by (1):

    (1)

    where c is the velocity of light, whereas the energy of the radiation is given by the Planck (2):

    (2)

    The latter equation indicates that frequency is a more fundamental quantity than wavelength, and many of the theoretical papers on spectroscopy use wave-numbers (σ = λ–1 with λ in cm) rather than wavelength. However, in correlative studies on natural products it is customary to express the absorption in terms of λ in millimicrons (mμ; 1 mμ = 10–7 cm), or less frequently in ångströms (Å; 1 Å = 10–8 cm). We have used mμ throughout the present work for such assignments. A selection of wavelengths, with corresponding wavenumber and energy values (from equations (1) and (2)) appear in Table 1. Since radiation of wave-

    TABLE 1

    WAVELENGTH-ENERGY RELATIONSHIPS

    length 200 mμ thus corresponds to an energy of ca. 143 kcal/mole, it is not surprising that the irradiation of organic structures should lead to such phenomena as trans cis isomerization, cleavage of C—C bonds, and generation of free radicals.

    Having defined the unit of wavelength as the millimicron (1 mμ), we can specify the limits of the electronic spectrum (30–300 kcal/(mole)) as extending from 800 to 100 mμ. Above 800 mμ we enter the near-infrared, while below 100 mμ lies the soft X-ray region. For physiological reasons, the visible region is defined as 400–800 mμ. The ultraviolet region is subdivided into the far ultraviolet (100–190 mμ) and near ultraviolet (190–400 mμ). Owing to the absorption of the shorter wavelengths by constituents of the atmosphere, it is necessary that measurements appreciably below 190 mμ should be carried out in vacuo. Experimentation in the far or vacuum ultraviolet region is difficult and much work in this field remains to be done. Our interest is mainly confined to the region 190–400 mμ, although, since there is no doubt that the essential process of energy absorption followed by excitation of electrons occurs in the 400–700 mμ region, we shall have occasion to discuss electronic absorption spectra in the visible region.

    In the cases of a few molecules where symmetry factors are favourable, it has been possible to arrive at theoretical prediction of wavelength of absorption from energy considerations and the use of equations (1) and (2). We shall, however, confine our discussion to the empirical approach throughout.

    Absorption Intensity

    The second fundamental characteristic required to define absorption of monochromatic light of wavelength λ, is the intensity of absorption. This is given by equation (3)

    (3)

    where I0 and I are the respective intensities of the incident and transmitted light and n is the number of molecules of absorbant in the light path. For substances in solution n is proportional to the molar concentration of solute (c) and to the length (l) of the cell that contains the solution. Transferring to decadic logarithms we have

    (4)

    Equation (4), which should be thought of as an experimental relationship, is described as the Beer–Lambert Law. The left side of (4) is variously described as the absorbance or optical density of the solution, symbolized by d. The commercial spectrophotometer provides a direct measurement of d either as a recorded trace or scale reading.

    is the molecular extinction coefficient is a pure number whose units are l mole–1cm–1 and from (4) we have

    (5)

    where c is in grams per litre. Where the molecular weight of the compound is unknown, the expression

    (sometimes called E value) is used where

    We have seen that the wavelength of absorption of a molecule is determined by the energy of the appropriate electronic transition. The extinction coefficient is controlled by the size of the absorbing species and the probability of the transition. It can be shown¹ that

    (6)

    where P is the transition probability and a ∼ 10⁵ for a transition of unit probability. The highest recorded values are in the order of cavalues of 10⁶ in the poly-yne series< 10³ low intensity absorptions. An absorption of 100–1000 corresponds to a transition of low probability (P = 0.01 or less) and such low intensity is sometimes termed a forbidden transition.

    Another fundamental value derivable from the tenets of electromagnetic theory is the oscillator strength (f) given by

    (7)

    where ⊿ν is the range of wavenumbers (reciprocal wavelengths) over which the electronic transition extends³ We may regard (f) as the measure of the number of electrons per molecule taking part in the transitions responsible for the absorption of light. We shall have occasion to refer to calculations using f values in Chapter 2.

    An empirical method of intensity calculation has been developed by Piatta for benzenoid compounds and was recently extended by Mason⁴b to the heteroaromatic series. Benzene has an absorption band near 260 mμ100), there being no change of dipole moment during the electronic transitions associated with this band. Replacement of a —CH= group by —N= or a hydrogen atom by a substituent, leads to perturbation and thence to dipole moment changes which are observed in the regular increase of the intensity of the 260 mμ band. For a given substituent, Platt has shown that equation (8) is valid:

    (8)

    where Ia is the intensity of the 260 mμ band for the substituted compound, I0 is that of the parent, benzene, and Ma is the spectroscopic moment of substituent (a), which is positive for electron-releasing (+M) and negative for electron-attracting (–M) substituents.

    Then for polysubstitution:

    (9)

    and

    (10)

    A set of Ma values has been evolved† and accounts qualitatively for the observed intensities of substituted benzenes and naphthalenes, the latter series making use of the 320 mμ band of naphthalene. Extension to pyridines and more complex N-heteroaromatic compounds has been described.⁴b

    It is now agreed that the Beer–Lambert Law is fundamentally true, although its application is not so general as was once believed. The law is, however, rigorously obeyed when a single species gives rise to the observed absorption. Deviations from the law may be observed:

    (i) when different forms of the absorbing system are in equilibrium; in this case it is often meaningful to measure the absorption at different values of pH. This results in a change of intensity for a given wavelength band, the change in ionic species being inversely proportional to the relative extinction coefficients. This means that there will be one point common to all the extinction curves which is constant for all values of pH. Such a point is called an isosbestic point and the presence of an isosbestic point for any given system is indicative that a number of equilibrium forms are present.

    (ii) When complexes occur between pairs of solute molecules (association) or between solute and solvent (e.g. H-bonding).

    (iii) When there is thermal equilibrium between the ground state and a lowlying electronically excited state, as in some forms of thermochromism.

    (iv) Beer’s Law is not obeyed by fluorescent compounds or compounds changed chemically by ultraviolet irradiation.

    Presentation of Absorption Spectra

    /wavelength (in mμ) with increasing wavelengths from left to right. However, as a result of publication policy and usage, it is probably true to say that less than 5% of ultraviolet spectra reported in the literature since 1930 have been presented in graphical form. A typical curve is illustrated in Fig. 1, and most investigators record their spectra in this way, although the results are published in the form: λmax (in mμ. Sometimes the minimum λmin is also reported. Projects in hand for the presentation of many reference spectra in graphical form⁵ should aid identification by inspection as many subtleties hitherto unsuspected might be revealed by graphical presentation of data – at the same time visual comparison would be facilitated. It is also frequently necessary to subtract and add absorption curves during the course of a structural investigation, so that although we shall concern ourselves in this volume mainly with the correlation of λmax and intensity values with structural features, the construction of an absorption curve is still a most important operation. The tedium of such a process is now largely removed by the advent of the commercial recording spectrophotometer.

    FIG. 1 Absorption Curve of benzoic acid in cyclohexane. (after Friedel and Orchin No. 71)

    Solvents

    In Fig. 2 we illustrate the effect of changing from hydrocarbon to hydroxylic solvent for the case of phenol. The loss of fine structure in the latter case is quite typical, the broad band due to H-bonded solvent–solute complexes replacing the fine structure present in hexane solution. The fine structure revealed in the latter solvent illustrates the principle that non-solvating or non-chelating solvents produce a spectrum near to that obtaining in the gaseous state. Common solvents and their window regions are shown in Table 2.

    TABLE 2

    MINIMUM WAVELENGTHS FOR COMMON SOLVENTS USED IN SPECTROSCOPY

    †Adopted throughout this work.

    FIG. 2 Absorption curve of phenol in: (a) iso-octane, (b) ethanol. (Coggleshall and Lavy, J.Amer.Chem.Soc, 70, 3283, 1948.)

    Experimental Methods and Instrumentation

    Detailed descriptions of the apparatus and technique of absorption spectrophotometry are so numerous that there is no need to mention them in this book except to say that the modern instrument is expected to be accurate to within 2 mμ in λ . Frequent calibration of the instrument is essential for accuracy of spectral correlations.

    Chromophores

    Although ultraviolet absorption marks the excitation of electrons from their ground state, the nuclei which the electrons hold together determine the strength of the binding. Thus, the characteristic energy of a transition, and hence the wavelength of absorption, is a property of a group of atoms rather than of electrons themselves. When such absorption occurs the group producing it is called a chromophore. We suggest that this general definition should be adopted in preference to other usages. Structural changes affecting such a group of atoms can be expected to modify its absorption. The way in which spectra may be used to derive structural information is the main theme of this book. It must be stated at the outset (spectroscopists need read no further) that at present the relationships we have chosen to discuss between electronic light absorption and molecular structure of complex molecules form empirical, but none the less useful working guides for the organic chemist. Some knowledge of the commonly occurring chromophoric systems is therefore an essential requirement of the practising natural product chemist.

    The three indispensable contributions to electronic absorption are the single bond (σ-electrons), the multiple bond (π-electrons) and the unshared electron pair (p-electrons)†. We shall not discuss free radical absorption (unpaired electrons) or ionic spectra (charge electrons).

    σ-Absorption

    Compounds containing only σ-valency electrons are the saturated hydrocarbons which absorb below 170 mμ, and give rise to intense bands corresponding to N V and N R transitions⁶,⁷. The latter type of transition is also called a Rydberg transition. In Table 3 will be found a representation of such absorptions together with the wavelengths at which they occur. These lie outside the range of the commercial instruments, the transparency of saturated hydrocarbons to 190 mμ making them ideal solvents for spectroscopy (Table 2).

    TABLE 3

    TYPES OF TRANSITION⁷a

    †The geometrical representation of the excited state in terms of molecular orbital treatment has been formulated for some simple molecules (ref. 13). The symbols in Table 3 are intended as convenient shorthand, pending a completely generalized notation.

    Absorption due to Electron Lone Pairs (p-electrons)

    The non-bonding p-electrons are held more loosely than the σ-electrons and undergo Rydberg (N R) transition at correspondingly higher wavelength. From equation (2) the ionization potentials shown in Table 4 have been determined spectroscopically. In addition to these bands in the far ultraviolet, atoms in the first row of the periodic table bound to carbon show a p-orbital electron transition to an antibonding σ-orbital (symbolized σ*), the transition being described as n σ* where n denotes the orbital of the non-bonding p-electron in the ground state. Some n σ* frequencies appear in Table 1.1 (p. 16).

    TABLE 4

    N → R TRANSITIONS AND IONIZATION POTENTIALS OF SATURATED COVALENT MOLECULES, R — X⁷

    π-electrons

    Unsaturated or multiple bond electrons present in olefinic, carbonyl or nitroso compounds for example, show Rydberg transitions in the 120–180 mμ region (10.5 eV). More importantly for diagnostic purposes, several of these chromophores have absorption in the more accessible regions of the spectrum. Thus olefins have intense bands between 180 and 205 mμ ∼ 10⁴) which are assigned to the transition of a π-electron from its orbital to an antibonding (π*) orbital, symbolized π π*∼ 15) near 280 mμ, corresponding to promotion of the non-bonding electron of oxygen to an antibonding π orbital (n π* band), a second band of high intensity being found near 190 mμ, ascribed to the nσ* transition.

    CONJUGATION OF CHROMOPHORES

    The presence of non-interacting chromophores in an organic molecule gives rise to absorption corresponding to the simple summation of the contributing absorptions. Conjugation of like or unlike chromophores gives rise to new and intense absorption bands in the 210–260 mμ region of the spectrum.

    π → π* Absorption

    Conjugation of like chromophores, as exemplified by butadiene,

    produces absorption at 217 mμ ∼ 10⁴) corresponding to the promotion of a π-electron to the antibonding π* orbital (π π* transition). Such absorption is affected by the further conjugation of p-electron systems with such a chromophore. Thus the presence of an atom X, with non-bonding electrons, can give rise to excited states represented for example

    Enjoying the preview?
    Page 1 of 1