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

Only $11.99/month after trial. Cancel anytime.

Introduction to Infrared and Raman Spectroscopy
Introduction to Infrared and Raman Spectroscopy
Introduction to Infrared and Raman Spectroscopy
Ebook801 pages8 hours

Introduction to Infrared and Raman Spectroscopy

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Now in its third edition, this classic text covers many aspects of infrared and Raman spectroscopy that are critical to the chemist doing structural or compositional analysis. This work includes practical and theoretical approaches to spectral interpretation as well as a discussion of experimental techniques. Emphasis is given to group frequencies, which are studied in detailed discussions, extensive tables, and over 600 carefully chosen and interpreted spectral examples. Also featured is a unique treatment of group frequencies that stresses their mechanical origin. This qualitative approach to vibrational analysis helps to simplify spectral interpretation.
Additional topics include basic instrumental components and sampling techniques, quantitative analysis, Raman polarization data, infrared gas contours, and polarized IR studies, among others.
  • Focuses on group frequency correlations and how to use them in spectral interpretation
  • Revised and updated by a pioneer in the field, Norman Colthup, who for thirty years has served as an expert lecturer for the Fisk Infrared Institute
  • Explores new group frequency studies in aromatics, alkanes and olefins, among others
  • Includes completely updated section on instrumentation
LanguageEnglish
Release dateOct 10, 1990
ISBN9780080917405
Introduction to Infrared and Raman Spectroscopy

Related to Introduction to Infrared and Raman Spectroscopy

Related ebooks

Science & Mathematics For You

View More

Related articles

Reviews for Introduction to Infrared and Raman Spectroscopy

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

    Introduction to Infrared and Raman Spectroscopy - Norman B. Colthup

    INTRODUCTION TO INFRARED AND RAMAN SPECTROSCOPY

    Third Edition

    Norman B. Colthup

    Lawrence H. Daly

    Stephen E. Wiberley

    Table of Contents

    Cover image

    Title page

    Copyright

    PREFACE

    Chapter 1: VIBRATIONAL AND ROTATIONAL SPECTRA

    Chapter 2: IR EXPERIMENTAL CONSIDERATIONS

    Chapter 3: MOLECULAR SYMMETRY

    Chapter 4: THE VIBRATIONAL ORIGIN OF GROUP FREQUENCIES

    Chapter 5: METHYL AND METHYLENE GROUPS

    Chapter 6: TRIPLE BONDS AND CUMULATED DOUBLE BONDS

    Chapter 7: OLEFIN GROUPS

    Chapter 8: AROMATIC AND HETEROAROMATIC RINGS

    Chapter 9: CARBONYL COMPOUNDS

    Chapter 10: ETHERS, ALCOHOLS, AND PHENOLS

    Chapter 11: AMINES, C=N, AND N=O COMPOUNDS

    Chapter 12: COMPOUNDS CONTAINING BORON, SILICON, PHOSPHORUS, SULFUR, OR HALOGEN

    Chapter 13: MAJOR SPECTRA–STRUCTURE CORRELATIONS BY SPECTRAL REGIONS

    Chapter 14: THE THEORETICAL ANALYSIS OF MOLECULAR VIBRATIONS

    INDEX

    Copyright

    Copyright © 1990 by Academic Press, Inc.

    All rights reserved.

    No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopy, recording, or any information storage and retrieval system, without permission in writing from the publisher.

    ACADEMIC PRESS, INC.

    1250 Sixth Avenue, San Diego, CA 92101

    United Kingdom Edition published by

    ACADEMIC PRESS LIMITED

    24-28 Oval Road, London NW1 7DX

    Cover design by Elizabeth E. Tustian

    Library of Congress Cataloging-in-Publication Data

    Colthup, Norman B.

    Introduction to Infrared and Raman spectroscopy / Norman B.

    Colthup, Lawrence H. Daly, Stephen E. Wiberley. — 3rd ed.

    p. cm.

    Includes bibliographical references.

    ISBN 0-12-182554-X (alk. paper)

    1. Chemistry, Organic. 2. Infrared spectroscopy. 3. Raman spectroscopy. I. Daly, Lawrence H. II. Wiberley, Stephen E. III. Title.

    QD272.S6C64 1990

    547.3′ 08583 — dc20 90-291

    CIP

    Printed in the United States of America

    90 91 92 93 9 8 7 6 5 4 3 2 1

    PREFACE

    In this third edition, the general plan of the previous editions has been retained in order to provide a book that covers in one volume those aspects of vibrational spectroscopy that a chemical spectroscopist will find useful in the study of chemical structure or composition. This includes introductory theory of vibrational and rotational spectra, basic infrared instrumental components and experimental techniques, quantitative analysis, the use of symmetry in vibrational spectroscopy, and a detailed example of theoretical vibrational analysis. The most extensive part of this book (Chapters 4–13) is an in-depth study of group frequency correlations and how to use them in spectral interpretation.

    Some years back, many chemists considered infrared spectroscopy to be one of their most useful tools and knew well how to apply it to problem solving. In recent years, other techniques, such as NMR and mass spectrometry, have captured the attention of chemists with the result that high-level infrared spectral interpretation is becoming something of a lost art. Infrared spectroscopy continues to be a powerful tool in the study of molecular structure or composition, but the full benefit of the technique cannot be derived without some knowledge of spectral interpretation: this is one of the main topics of this book. Our understanding of the nature of functional group vibrations continues to grow, and important new contributions have been added in this edition.

    In Chapter 1 we discuss some details of how molecular vibrations and rotations interact with radiation to create the infrared and Raman spectra. The theoretical aspects are presented on an introductory level. While there are no detailed quantum mechanical calculations provided, infrared and Raman spectra are discussed from the quantum viewpoint. Classical analogies, which are usually easier to understand, are often referred to for clarification. Diatomic vibrations and diatomic vibrational-rotational infrared spectra are discussed in some detail. These serve as models for the spectra of larger molecules of more general interest. In the present edition, we have made modifications intended to clarify some of the explanations of the vibration-radiation interactions.

    The biggest changes that have occurred in recent years have been in instrumentation, and the chapter on IR experimental techniques has been revised extensively. We have avoided detailed descriptions of instruments; rather we have described in broad terms the basic features of instrumentation. These include discussions of interferometers that are used in Fourier transform IR spectrometers, and discussions of grating monochromators that are used in dispersive IR spectrometers. Infrared spectra can be run on gases, liquids, or solids; most of the techniques used in running these samples are discussed. Solid sampling techniques are the most diversified. Sections are included on quantitative analysis, internal reflection spectroscopy, and the use of polarized IR radiation.

    If a molecule has some symmetry, this may modify features of the infrared and Raman spectra. These symmetry effects can be helpful in the interpretations of the spectra; they are reviewed in Chapter 3.

    In Chapter 4, we discuss the vibrational origin of group frequencies, that is, infrared frequencies that are absorbed by certain chemical functional groups of atoms, such as carbonyls. Mathematical procedures such as those outlined in Chapter 14 are available for the rigorous calculation of molecular vibrational frequencies from a given set of force constants, but are impractical for use in day-to-day chemical structure problems. There is a need for qualitative approximate methods for analyzing the nature of functional group vibrations to give meaning to simple spectra-structure correlations derived empirically. In this chapter, many applications of these qualitative principles are included, and a new application has been added on some substituted benzene vibrations.

    In Chapters 5–12, individual functional groups and their associated group frequencies are discussed in detail. These include such groups as CH2, CH3, X≡ Y, X=Y=Z, C=C, aromatic rings, C=O, C=O, OH, NH, C=N, NO2, and groups containing B, Si, P, S, or halogen. Group frequencies are discussed and referenced in the text, are summarized in numerous tables, and are illustrated in many cases. In the present edition, additions have been made to almost every chapter, most notably on methyl and methylene groups, olefin groups, and aromatic groups. A number of new, simple calculations of group frequency shifts have been added, and many new illustrations are included that show generalized spectra of various groups. Throughout these chapters, the vibrational origin of the group frequencies is emphasized since, ultimately, this makes the spectra easier to interpret. As in earlier editions, we discuss mainly infrared group frequencies, but some Raman group frequencies have been included, especially when they supplement the IR study.

    In Chapter 13, group frequencies are considered in terms of the spectral regions in which they occur. The first part of the chapter outlines an orderly procedure for the initial interpretation of an unknown IR spectrum by regions. Then, spectra–structure correlations are shown in chart form where one can look for groups that absorb in a given region, or look for regions where a given group absorbs. The most important section includes 624 interpreted IR spectra arranged by groups, along with 36 Raman spectra. This book is not a library, but its purpose is to provide interpreted spectral examples of functional group frequencies. This is one of the best ways of verifying a proposed assignment.

    Since the full understanding of molecular vibrations is based on vibrational analysis of small molecules, it is helpful to have some intermediate level understanding of the procedures involved. A spectroscopist can then better understand the extensive literature on the subject. This approach is not new, and easy access to computers has changed the way calculations are actually done; however, the emphasis here lies in developing the various required steps in an unusually detailed manner. We devoted considerable effort to presenting the analysis so that the average student will follow each step easily. While some use is made of group theory and matrix algebra, no detailed previous knowledge of these is necessary since sufficient material is included within this text. The purpose of this third edition is not to present a complete coverage of all the ramifications of the normal coordinate methods but, rather, to serve as a guide to the beginner for the procedures that are covered.

    Norman B. Colthup, Lawrence H. Daly and Stephen E. Wiberley

    CHAPTER 1

    VIBRATIONAL AND ROTATIONAL SPECTRA

    Publisher Summary

    The energy of a molecule consists partly of translational energy, rotational energy, vibrational energy, and electronic energy. Electronic energy transitions give rise to the absorption or emission in the ultraviolet and the visible regions of the electromagnetic spectrum. Pure rotation gives rise to the absorption in the microwave region or the far infrared. Molecular vibrations give rise to the absorption bands throughout most of the infrared region of the spectrum. This chapter focuses on vibrational and rotational spectra and discusses how molecular vibrations and rotations interact with radiation to create the infrared and Raman spectra. The vibrational and the rotational frequencies of molecules can be studied by Raman spectroscopy and by infrared spectroscopy. While they are related to each other, the two types of spectra are not exact duplicates and each has its individual strong points. In Raman spectroscopy, only the wave number is used. Infrared and Raman spectrum both involve vibrational and rotational energy levels; they are not duplicates of each other but rather complement each other. This is because the intensity of the spectral band depends on how effectively the photon energy is transferred to the molecule.

    1.1 Introduction

    The energy of a molecule consists partly of translational energy, partly of rotational energy, partly of vibrational energy, and partly of electronic energy. For a first approximation these energy contributions can be considered separately. Electronic energy transitions normally give rise to absorption or emission in the ultraviolet and visible regions of the electromagnetic spectrum. Pure rotation gives rise to absorption in the microwave region or the far infrared. Molecular vibrations give rise to absorption bands throughout most of the infrared region of the spectrum. In this book we shall mainly be concerned with the interaction of electromagnetic radiation with molecular vibrations and rotations.

    Electromagnetic radiation is characterized by its wavelength λ (the length of one wave), its frequency v , in waves per centimeter (cm−1), is related to the other parameters by

    (1.1)

    where c is the velocity of light in a vacuum (2.997925 × 10¹⁰ cm/sec), and (c/n) is the velocity of light in a medium whose refractive index is n, in which the wavenumber is measured. The refractive index of air is 1.0003. The frequency v is independent of the medium and is expressed in cycles per second (sec−1) or Hertz (Hz), and v is the wavelength in cm. In terms of these units, the ultraviolet, visible, infrared, and microwave regions of the electromagnetic spectrum assume the values shown in Table 1.1.

    TABLE 1.1

    VALUES FOR λ, , AND λ

    In the infrared region of the electromagnetic spectrum the practical unit for wavelength is 10−4 cm or 10−6 m. This unit has long been called the micron, μ, but is now called the micrometer, μm. Wavelength is a property of radiation but not a property of molecules. The properties that radiation and molecules have in common are energy and frequency. The frequency in Hz in the infrared part of the spectrum is an inconveniently large number and the wavenumber (cm−1) which is proportional to frequency [see Eq. (1.1)] is more commonly used. Infrared spectra have been presented so that the horizontal coordinate is either linear with wavelength (v) or with wavenumber (cm−1), with usually a factor of two scale expansion below 2000 cm−1. The wavenumber is sometimes referred to as the frequency in cm−1. The term frequency in cm−1 is understood to mean the frequency (in Hz) divided by the velocity of light (in cm/sec) or the reciprocal of the radiation wavelength (in cm).

    The vibrational and rotational frequencies of molecules can be studied by Raman spectroscopy as well as by infrared spectroscopy. While they are related to each other, the two types of spectra are not exact duplicates and each has its individual strong points. In Raman spectroscopy, only the wavenumber is used.

    1.2 Photon Energy

    According to the quantum theory the energy of a photon Ep is given by

    (1.2)

    where h is Planck’s constant, 6.6256 × 10−27 erg-sec, or 6.6256 × 10−34 joule-sec. This photon energy may be absorbed or emitted by a molecule in which case the rotational, vibrational, or electronic energy of the molecule will be changed by an amount ΔEm. According to the principle of conservation of energy

    (1.3)

    If the molecule gains energy, ΔEm is positive and a photon is absorbed. If the molecule loses energy, ΔEm is negative and a photon is emitted. A commonly used form of Eq. (1.3) is

    (1.4)

    which reads: The wavenumber of the absorbed or emitted photon is equal to the change in the molecular energy term expressed in cm−1 (E/hc).

    1.3 Degrees of Freedom of Molecular Motion

    In the study of molecular vibrations we can start with a classical model of the molecule where the nuclei are represented by mathematical points with mass. The internuclear forces holding the molecule together are assumed to be similar to those exerted by massless springs which tend to restore bond lengths or bond angles to certain equilibrium values. Each mass requires three coordinates to define its position, such as x, y, and z in a cartesian coordinate system for example. As a result it has three independent degrees of freedom of motion, in the x, y, or z direction. If there are N atomic nuclei in the molecule, there will be a total of 3N degrees of freedom of motion for all the nuclear masses in the molecule.

    The center of gravity of the molecule requires three coordinates to define its position and therefore has three independent degrees of freedom of motion which are translations of the center of gravity of the molecule. When a nonlinear molecule is in its equilibrium configuration, it requires three rotational coordinates to specify the molecular orientation about the center of gravity. For example, these can be three angular coordinates specifying rotation about three mutually perpendicular axes, each going through the center of gravity. A nonlinear molecule, therefore, has three independent rotational degrees of freedom. A linear molecule only has two independent rotational degrees of freedom about two mutually perpendicular axes, perpendicular to the molecular axis. Rotation of a linear molecule about the molecular axis is not considered a degree of freedom of motion since no displacements of nuclei are involved.

    After subtracting the translational and rotational degrees of freedom from the total 3N degrees of freedom, we are left with 3N – 6 internal degrees of freedom for a nonlinear molecule and 3N – 5 internal degrees of freedom for a linear molecule. Translations of the center of gravity and rotations about the center of gravity can all take place independently without any change occurring in the shape of the molecule without moving the center of gravity and without rotating the molecule.

    1.4 Normal Modes of Vibration

    It can be shown (Chapter 14) that the 3N − 6 internal degrees of freedom of motion of a nonlinear molecule correspond to 3N − 6 independent normal modes of vibration. In each normal mode of vibration all the atoms in the molecule vibrate with the same frequency and all atoms pass through their equilibrium positions simultaneously. The relative vibrational amplitudes of the individual atoms may be different in magnitude and direction but the center of gravity does not move and the molecule does not rotate. If the forces holding the molecule together are linear functions of the displacement of the nuclei from their equilibrium configurations, then the molecular vibrations will be harmonic. In this case each cartesian coordinate of each atom plotted as a function of time will be a sine or cosine wave when the molecule performs one normal mode of vibration (see Fig. 1.1).

    FIG. 1.1 Normal mode of vibration of a ball and spring model of a diatomic molecule such as HCl. The displacement versus time plot for each mass is a sine wave and the center of gravity (dashed line) is motionless.

    1.5 Mechanical Molecular Models

    Normal modes of vibration can be demonstrated experimentally using vibrating mechanical models.¹ One such arrangement is illustrated in Fig. 1.2. A source of oscillation is provided by an eccentric shaft on the side of the main shaft of a motor whose rotational speed is adjustable. A ball and spring model of a molecule such as CO2 is suspended by long threads, one attached to each ball. This leaves the model free to move in the horizontal plane. A short length of thread is used to connect the eccentric to the suspension thread of an end ball about 6 in. above the ball. This horizontal coupling

    FIG. 1.2 Demonstration of molecular vibrations using ball and spring molecular models, in this case of a molecule such as CO2.

    thread is held at an angle to the model axis so that as the motor rotates the eccentric can exert weak periodic forces on the model, both along and across the model axis. As the motor speed is varied, the frequency of the disturbing force is changed. When the disturbing force frequency matches one of the natural frequencies of vibration of the model, resonance occurs. The model responds by performing one of its normal modes of vibration (see Fig. 1.2) where it can be seen that all the masses perform simple harmonic motions with the same frequency and go through their equilibrium positions simultaneously. In some modes some masses may vibrate with zero amplitude, that is, they may remain motionless. At a different disturbing force frequency another normal mode of vibration may be activated. When the disturbing force frequency does not coincide with the frequency of a normal mode of vibration, the model remains nearly motionless.

    Mechanical molecular models have their limitations because the masses are not mathematical points and the helical springs are not massless and are only an approximate representation of the molecular force field. For demonstration purposes, however, the relative frequencies obtained from a properly constructed model agree sufficiently well with the relative frequencies of the actual molecule. The ratio of the stretching to bending force constants of the spring should be adjusted so that the ratio is about 8 or 10 to 1 which is approximately the ratio found in real molecules. This ratio can be increased if too small by permanently stretching out the spring a little. The model specifications in Fig. 1.2 are not critical. If larger masses are used, stiffer springs can be used. However, for clarity of visual observation, the frequencies should be in the 1–7 cycles/sec range.

    If, instead of being excited with the motor, the model is simply struck with a hammer, the resulting motions appear to be erratic and nonsinusoidal. However, closer inspection will show that the motions consist of a simultaneous performance of all the normal modes of vibrations excited previously plus translations and rotations. In the case of the model in Fig. 1.2, free translations and rotations become pendulum oscillations of the model as a whole because of the suspension system, but these oscillations are made considerably lower in frequency than the lowest vibrational frequency and therefore cause no difficulties.

    The linear CO2 molecule has 3N − 5 or 4 fundamental modes of vibration. The model discussed here has only three because it is constrained by the suspension system to move in a plane. The fourth fundamental is an out-of-plane bending vibration identical in character and frequency to the in-plane bending vibration performed by the model. Such pairs of vibrations which have the same frequency are called doubly degenerate vibrations. A bending vibration in any direction perpendicular to the molecular axis can be described as a linear combination of the two fundamental components.

    1.6 Coordinates Used to Describe Molecular Vibrations

    The stretching vibrations of the model in Fig. 1.2 representing a molecule like CO2 are illustrated in Fig. 1.3 along with several types of coordinates used for molecular vibrations. Cartesian displacement coordinates X measure displacement from the equilibrium position of each nuclear mass using cartesian coordinates. Each atom has its own cartesian coordinate system with the origin defined by the equilibrium position of the atom. Internal coordinates measure the change in shape of a molecule as compared to its equilibrium shape without regard for the molecules’ position or orientation in space. For example, there can be a change in a bond length or a bond angle from its equilibrium value. In Fig. 1.3 the bond length internal coordinates are represented by S1 and S2, where S1 is one bond length at any time minus the equilibrium bond length, and S2 is similarly defined.

    FIG. 1.3 The stretching vibrations of a linear triatomic model and types of coordinates used to characterize the vibrations. The cartesian coordinates X are mass displacements. The internal coordinates S are bond length changes. In this case S1 = X1 – X2 and S2 = X2 – X3. During the performance of one normal mode of vibration, these coordinates have values in specified proportions to the value of one oscillating normal coordinate Q.

    Normal modes of vibration may be described using either cartesian displacement coordinates or internal coordinates. For every normal mode of vibration all the coordinates vary periodically with the same frequency and go through equilibrium simultaneously. The form of the normal mode of vibration is defined by specifying the relative amplitudes (which may be positive or negative) of the various coordinates in the set being used. For example, in Fig. 1.3, S1/S2 equals minus one and one for the two vibrations, respectively.

    Each normal mode of vibration can also be characterized by a single normal coordinate Q, which varies periodically. A given normal coordinate is a measure of the amplitude of a specific normal mode of vibration. As one normal coordinate vibrates, every cartesian displacement coordinate and every internal coordinate vibrates, each with an amplitude in specified proportion (positive or negative) to the amplitude of the normal coordinate, the proportions being such that the resulting motion is a normal mode of vibration. Since normal modes of vibration can be excited independently of each other, normal coordinates are also independent in the sense that each normal coordinate makes a separate and independent contribution to the total vibrational potential and kinetic energy. This simplifies the quantum mechanical formulation where contributions from each normal coordinate can be treated separately.

    1.7 Classical Vibrational Frequency Formula for a Diatomic Molecule

    Because of its importance in the study of molecular vibrations, the classical vibrational frequency for a diatomic molecule will be derived. Let the diatomic molecule be represented by two masses m1 and m2 connected by a massless spring (see Fig. 1.4). For simplicity the masses may be allowed to move only along the molecular axis, like beads on a wire, for example. Let the displacement of each mass from equilibrium along the axis be X1 and X2. In this case (X2 – X1) is the amount the bond length differs from the equilibrium length. It will be assumed that the bond spring obeys Hooke’s law which means that if the bond is not at its equilibrium length, each mass will experience a force equal to a constant, F, times the bond length distortion from equilibrium (X2 – X1). We can equate the Hooke’s law force, exerted along the spring axis on each mass, to mass m times acceleration (d²X/dt²) along this same axis (Newton’s law) to obtain two equations, one for each mass.

    FIG. 1.4 Diatomic molecular model. On the right are the two solutions to the equations of axial motion where mass displacement as a function of time is plotted for translation and vibration.

    (1.5)

    When (X2 – X1) is positive (stretched spring) the force on mass m1 is directed in the plus X direction and the force on m2 is directed in the minus X direction which accounts for the negative sign in the m2 equation.

    Since these equations of motion are a function of X and its second derivative, one possible solution is that both X1 and X2 vary periodically as a sine or cosine function of time (ωt), where ω is a constant, since the second derivative of either function with respect to time is proportional to the original function.

    The period of the cosine function (in seconds per cycle) is equal to the reciprocal of the frequency v (in cycles per second). If we increase the time t by one period (1/v) from some initial time t0 to (t0 + 1/v), the cosine function for X will go through one cycle and repeat itself. Also after one cycle, the angle in the cosine function will increase by 2π radians. Therefore, after one cycle,

    By equating the two angle functions above, it can be seen that the constant ω in the cosine function is equal to 2πv. For the most general case we must add a constant A, to define the maximum amplitude of the cosine function, and a phase constant α, to define the cosine angle when t = 0. These are determined by the initial motions and displacements of the masses. With these additions we can write two cosine functions as trial solutions for X1 and X2 in the equations of motion in Eq. (1.5) as

    (1.6)

    Note that we are assuming at this stage that v and α have the same values for X1 and X2. Shortly, it will be clear that, if this is not true, then Eq. (1.6) will not be a solution to Eq. (1.5). The trial solutions for X1 and X2 in Eq. (1.6) describe simple harmonic motions for both masses, each oscillating as a cosine function of time t. The frequency v is in cycles per second, 2πv has the units of angular velocity in radians per second, and 2πvt and α are in radians. We note that while each mass oscillates along the spring axis with a different maximum amplitude A1 and A2, each mass oscillates with the same frequency v and phase constant α, which means that both masses go through their equilibrium positions simultaneously. Differentiating these equations twice with respect to time t we obtain

    and

    (1.7)

    Substituting the values for X1 and X2 [Eq. (1.6)] and their second derivatives [Eq. (1.7)] into the original equations of motion [Eq. (1.5)] we obtain, after cancelling the common factor cos(2πvt + α),

    (1.8)

    If v and α did not have the same values for X1 and X2 in Eq. (1.6), the common factor cos (2πvt + α) for X1 and X2 would now be different and could no longer be cancelled. These equations can be rearranged to give the amplitude ratios A1/A2

    (1.9)

    If v had different values for X1 and X2 in Eq. (1.6), then v in the m1 and m2 parts of Eq. (1.9) would have different values and a solution could not be obtained. Eliminating A1/A2 in Eq. (1.9) and solving for v we obtain.

    so

    (1.10)

    In one solution we obtain v = 0 and substituting this into either amplitude equation [Eq. (1.9)] we obtain A1 = A2. For any time t [from Eq. (1.6)] we obtain X1 = X2. This nongenuine zero frequency vibration is actually a translation in the X or spring axis direction. For the other solution we obtain

    (1.11)

    where we note the vibrational frequency is independent of amplitude. When this value of v is substituted into either amplitude equation [Eq. (1.9)] we obtain

    (1.12)

    or at any time t [from Eq. (1.6)] X1/X2 = –m2/m1 which indicates that the center of gravity does not move. These two solutions for v correspond to the two degrees of freedom the two masses have in the spring axis direction, (1) translation without vibration and (2) vibration without translation, (see Fig. 1.4). The trial solution in Eq. (1.6) will be a true solution for Eq. (1.5) if v is replaced by either of its equivalent values found above, because with this substitution for v the calculated result in Eq. (1.10) proves to be an equality.

    Sometimes Eq. (1.11) is written using the reduced mass u as

    (1.13)

    where

    Given the frequency and the reduced mass, the force constant can be calculated by

    (1.14)

    In Table 1.2 we present some calculated force constants for some diatomic molecules. These are arranged in order of decreasing force constants including triple bonds (N≡N), double bonds (O=O), single bonds (H—Cl) and ionic bonds (Na+Cl−). Reduced masses are given in unified atomic mass units (u) and kilograms (kg) and force constants are given in newtons per meter (N m−1). In the cgs system, force constants are given in dynes per centimeter. One newton per meter equals 10³ dyne cm−1. Molecular force constants are often expressed in millidynes per angstrom. One newton per meter equals 10−2 mdyne Å−1 and 1 dyne cm−1 = 10−5 mdyne Å−1.

    TABLE 1.2

    OBSERVED VIBRATIONAL FREQUENCIES AND CALCULATED FORCE CONSTANTS FOR SELECTED DIATOMIC MOLECULES

    1.8 Infrared Absorption and the Change in Dipole Moment

    In the mechanical analogy illustrated in Fig. 1.2 the model analogous to the molecule has certain natural vibrational frequencies. The eccentric on the motor provides an oscillating disturbing force analogous to infrared radiation oscillations which exert forces on the molecule. When the disturbing frequency matches the natural vibrational frequency of the model, the model absorbs energy from the motor and thereby increases its own vibrational energy by vibrating with increased amplitude. At nonresonant frequencies the model is ineffective in absorbing energy from the motor. This is analogous to infrared absorption where the molecule absorbs radiation energy by increasing its own vibrational energy. In a spectrometer the molecule is irradiated with a whole range of infrared frequencies but is only capable of absorbing radiation energy at certain specific frequencies which match the natural vibrational frequencies of the molecule (see Fig. 1.5), and these occur in the infrared region of the electromagnetic spectrum.

    FIG. 1.5 Schematic illustration of infrared absorption. The HCl molecule whose vibrational frequency is 8.67 × 10¹³ Hz increases its vibrational energy by absorbing the energy of an infrared photon which has this same frequency.

    While the absorption frequency depends on the molecular vibrational frequency, the absorption intensity depends on how effectively the infrared photon energy can be transferred to the molecule, and this depends on the change in the dipole moment that occurs as a result of molecular vibration. The dipole moment is defined, in the case of a simple dipole (see Fig. 1.6), as the magnitude of either charge in the dipole multiplied by the charge spacing. In a complex molecule this simple picture can be retained if the positive particle represents the total positive charge of the protons concentrated at the center of charge of the protons, and the negative particle represents the total negative charge of the electrons concentrated at the center of charge of the electrons. The center of charge of the protons coincides with the center of gravity of the protons, and the center of charge of the electrons coincides with the center of gravity of the electrons.

    FIG. 1.6 Forces generated on a dipole by an oscillating electric field. These forces tend to alternately increase and decrease the dipole spacing.

    Since the wavelength of infrared radiation is far greater than the size of most molecules, the electric field of the photon in the vicinity of a molecule can be considered uniform over the whole molecule. The electric field of the photon exerts forces on the molecular charges and by definition the forces on opposite charges will be exerted in opposite directions (see Fig. 1.6). Therefore the oscillating electric field of the photon will exert forces tending to change the spacing between the proton and electron centers of charge, thereby tending to induce the dipole moment of the molecule to oscillate at the frequency of the photon.

    At certain frequencies a forced dipole moment oscillation will tend to activate a nuclear vibration. These are molecular vibrational frequencies where the nuclear vibration causes a change in dipole moment. The more the dipole moment changes during a vibration, the more easily the photon electric field can activate that vibration. If a molecular vibration should cause no change in dipole moment then a forced dipole moment oscillation cannot activate that vibration. This is summarized in the selection rule that in order to absorb infrared radiation, a molecular vibration must cause a change in the dipole moment of the molecule. It can be shown that the intensity of an infrared absorption band is proportional to the square of the change in dipole moment, with respect to the change in the normal coordinate (∂μ/Q)² of the molecular vibration, giving rise to the absorption band.

    An alternate and easier way to picture the dipole moment is to consider the atoms in the molecule as charged particles rather than the individual electrons and protons. If an atom has the same number of electrons and protons it is electrically neutral and does not contribute to the dipole moment. However, well understood chemical forces in the molecule tend to redistribute the electrons somewhat, so that a given atom may have a little electron excess or deficiency, and may be considered as a particle with a small negative or positive charge. These charges may be derived, for example, from a molecular orbital calculation. A complication is that, unlike the fixed charge on the proton or electron, the charge on an atom may change during a molecular vibration, because the molecular geometry change may affect the electron redistribution. In any case, the positive particle of the molecular dipole represents the total excess positive charge on the atoms, concentrated at the center of the excess positive charge, and the negative particle of the dipole represents the total excess negative charge on the atoms, concentrated at the center of the excess negative charge.

    In the HCl molecule, the electronegative chlorine atom will have a slight excess of negative charge and the hydrogen atom will have a slight excess of positive charge. These two atoms can be looked upon as constituents of a simple dipole [see Fig. 1.7(a)]. During the vibration of the HCl molecule, the dipole spacing changes and the excess charge distribution also changes somewhat. Both of these effects change the dipole moment which means that this vibration is infrared active. The oscillating electric field of the photon will exert forces on the positively charged atom in one direction and on the negatively charged atom in the opposite direction. This will tend to excite the molecular vibration if the photon frequency matches the vibration frequency.

    FIG. 1.7 Dipole moment changes in certain molecular vibrations.

    In the hydrogen molecule, H2 in Fig. 1.7(b), a center of symmetry is present. As seen in Fig. 1.7(c), this means that, at equilibrium, every atom which is not at the center of the molecule, has an exactly equivalent atom on the direct opposite side of the center, and equidistant from it. This arrangement guarantees that the center of positive charge and the center of negative charge will coincide at the center of the molecule, which specifies a dipole moment of zero. During the H2 vibration in Fig. 1.7(b) the vibrationally-distorted molecule retains the center of symmetry so the dipole moment does not change, since it remains zero. This vibration is infrared inactive since there is no vibrationally caused change in dipole moment with which the electric field of the photon can interact. The electric field of the photon cannot pull the two equivalently charged hydrogen atoms in opposite directions as required in a diatomic vibration. This type of vibration is active in the Raman effect to be discussed later.

    In the CO2 molecule, a center of symmetry is present at equilibrium [Fig. 1.7(g)] which is retained during the in-phase stretching vibration in Fig. 1.7(f). This vibration is symmetric with respect to the center of symmetry and is infrared inactive. The electric field of the photon cannot pull the two equivalently charged oxygens in opposite directions as required in this vibration.

    FIG. 1.17 The radiation’s electric field interacting with the rotating dipole moment of a rotating diatomic molecule.

    A vibrationally distorted polyatomic molecule may have less symmetry than the molecule in the equilibrium configuration, as seen in the other two CO2 vibrations, in Fig. 1.7(d) and (e). These vibrationally distorted molecules do not have a center of symmetry. Such vibrations are antisymmetric with respect to the center of symmetry and may be infrared active. If the center of negative charge is assumed to be halfway between the two electronegative oxygen atoms and the center of positive charge is assumed to be at the carbon atom, then it can be seen in Fig. 1.7(d) and (e), that both the antisymmetrical, out-of-phase stretching vibration and the bending vibration cause a change in the dipole moment along or across the molecular axis, and so both are infrared active. The electric field of the photon pulls the positively charged carbon atom in one direction and the negatively charged oxygen atoms in the opposite direction along or across the molecular axis as required to excite these vibrations.

    It should be clear from the above arguments that, if a molecule has a center of symmetry at equilibrium, then vibrations during which the center of symmetry is retained will be infrared inactive. However, molecules that do not have a center of symmetry may also have infrared inactive vibrations if other types of symmetry such as planes or axes of symmetry are present as discussed in Chapter 3. An example is seen in Fig. 17(i) which shows the in-phase stretch of the carbonate ion (CO3)²–. This ion does not have a center of symmetry but during this vibration, the center of negative charge from the three negatively charged oxygens always coincides with the center of positive charge at the carbon, making this vibration infrared inactive. In Fig. 17(h) we show only one of the infrared active vibrations of (CO3)²–. This one involves out-of-phase CO3 stretch and shows some similarity to the out-of-phase CO2 stretch in Fig. 17(d).

    1.9 Anharmonicity and Overtones

    Up to now we have been discussing only harmonic vibrations. Mechanical anharmonicity results if the restoring force is not linearly proportional to the nuclear displacement coordinate. Electrical anharmonicity results

    Enjoying the preview?
    Page 1 of 1