You are on page 1of 1049
Handbook of Raman Spectroscopy 1. 12. 13. 14. 15. 16. 17. 18. 19. 20. 24 22. 23, 24, 25. 27. 28. A SERIES Infrared and Raman Spectroscopy (in three parts), edited by Edward G. Brame, Jr, and Jeanette G. Grasselli X-Ray Spectrometry, edited by H. K. Herglotz and L. S. Birks Mass Spectrometry (in two parts), edited by Charles Merritt, Jr, and Charles N. McEwen Infrared and Raman Spectroscopy of Polymers, H. W. Siesler and K. Holland-Moritz NMR Spectroscopy Techniques, edited by Ceci! Dybowski and Robert L. Lichter Infrared Microspectroscopy: Theory and Applications, edifed by Robert G. Messer- schmidt and Matthew A. Harthcock Flow Injection Atomic Spectroscopy, edited by Jose Luis Burguera Mass Spectrometry of Biological Materials, edited by Charles N. McEwen and Barbara S. Larsen Field Desorption Mass Spectrometry, Laszi6 Prokai . Chromatography/Fourier Transform Infrared Spectroscopy and Its Applications, Robert White Modern NMR Techniques and Their Application in Chemistry, edited by Alexander |. Popov and Klaas Hallenga Luminescence Techniques in Chemical and Biochemical Anal G. Baeyens, Denis De Keukeleire, and Katherine Korkiais Handbook of Near-Infrared Analysis, edited by Donald A. Burns and Emil W. Ciurczak Handbook of X-Ray Spectrometry: Methods and Techniques, edited by René E. Van Grieken and Andrzej A. Markowicz Internal Reflection Spectroscopy: Theory and Applications, edited by Francis M. Mirabella, Jr. roscopic and Spectroscopic Imaging of the Chemical State, edited by Michael D. Morris Mathematical Analysis of Spectral Orthogonality, John H. Kalivas and Patrick M, Lang Laser Spectroscopy: Techniques and Applications, E. Roland Menzel Practical Guide to Infrared Microspectroscopy, edited by Howard J. Humecki Quantitative X-ray Spectrometry: Second Edition, Ron Jenkins, R. W. Gould, and Dale Gedcke NMR Spectroscopy Techniques: Second Edition, Revised and Expanded, edited by Martha D. Bruch Spectrophotometric Reactions, Irena Nemcova, Ludmila Cermakova, and Jiri Gasparic Inorganic Mass Spectrometry: Fundamentals and Applications, edited by Christopher M. Barshick, Douglas C. Duckworth, and David H. Smith Infrared and Raman Spectroscopy of Biological Materials, edited by Hans-Ulrich Gremilich and Bing Yan Near-Infrared Applications In Biotechnology, edited by Ramesh Raghavachari edited by Willy R. . Ultrafast Infrared and Ramen Spectroscopy, edited by M. D. Fayer Handbook of Near-Infrared Analysis: Second Edition, Revised and Expanded, edited by Donald A. Bums and Emil W. Ciurczak Handbook of Raman Spectroscopy: From the Research Laboratory to the Process Line, edited by lan R. Lewis and Howell G. M. Edwards Copyright © 2001 by Taylor & Francis Group, LLC Applied Electrospray Mass Spectrometery, edited by Ashit K. Ganguly, Birendra Pra- ‘manik, and Michael L. Gross Ultraviolet Spectroscopy and UV Lasers, edited by Prabhakar Misra and Mark A Dubinskii Near-Infrared Spectroscopy in Pharmaceutical Applications, Emil W. Ciurezak and James Drennen Copyright © 2001 by Taylor & Franeis Group, LLC Handbook of Raman Spectroscopy From the Research Laboratory to the Process Line edited by lan R. Lewis Kaiser Optical Systems, Inc. Ann Arbor, Michigan Howell G. M. Edwards University of Bradford Bradford, West Yorkshire, England Manrcet Dekker, INc. New York + Base. Copyright © 2001 by Taylor & Franeis Group, LLC ISBN: 0-8247-0557-2 This book is printed on acid-free paper. Headquarters Marcel Dekker, Inc. 270 Madison Avenue, New York, NY 10016 tel: 212-696-9000; fax: 212-685-4540 Eastern Hemisphere Distribution Marcel Dekker AG Hutgasse 4, Postfach 812, CH-4001 Basel, Switzerland tel: 41-61-261-8482; fax: 41-61-261-8896 World Wide Web http:/www.dekker.com The publisher offers discounts on this book when ordered in bulk quantities. For more information, write to Special Sales/Professional Marketing at the headquarters address above. Copyright © 2001 by Marcel Dekker, Inc. All Rights Reserved. Neither this book nor any part may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopying, microfilming, and recording, or by any information storage and retrieval system, without permission in writing from the publisher. Current printing (last digit) 09876543 PRINTED IN THE UNITED STATES OF AMERICA 2001 by Taylor & Francis Group, LLC Ian Lewis would like to thank his mother, Valerie, his wife, Mary, and his children, Valerie, Henry, Maria, and Edward, for their unceasing support during this project. In addition he would like to thank Mary for proofreading much of this work and for her help in compiling the index. Howell Edwards dedicates his contributions to this volume to his wife, Gill, and daughter, Kate, for their continued support and encouragement over many years, and to the memory of Dr. Leonard Woodward, who first introduced him to Raman Spectroscopy as a research student, which developed into a lifelong study. Finally, both editors express their gratitude to all the chapter authors who spent many hours on their respective contributions and met many deadlines despite their very busy schedules. Copyright © 2001 by Taylor & Franeis Group, LLC Preface Since 1928, when Raman and Krishnan published their manuscript in Nature that heralded the experimental discovery of a “new type of secondary radiation,” —later to be termed Raman scattering—the field has undergone several distinct phases. From its discovery until World War II, when a commercial infrared spectrometer was produced, Raman spec- troscopy retained its position as the predominant vibrational spectroscopic technique. Since the widespread adoption of infrared spectroscopy, Raman has existed as a poor cousin but has had two notable periods of user expansion. The first, in the 1960s, coincided with the development of the double monochromator, the laser, and electronic methods of signal detection. The second and current period started approximately 15 years ago and is still progressing. During this time, three major developments have driven the field of Raman spectroscopy forward: the advent of laser-line blocking filters, the successful demonstration and quick acceptance of FT-Raman spectroscopy, and the adoption of scientific-grade charge coupled device (array) detectors as the preferred detector for the majority of dis- persive Raman spectroscopic experiments. The developments noted above have indeed changed the perception of Raman spec- troscopy completely. However, it should be noted that there are still several applications in which using a triple monochromator equipped with either a PMT or an array detector may be the only appropriate technology. This is especially true for applications in which low-frequency vibrations need to be observed. ‘These instrumental developments have had a dramatic effect on the applications of Raman spectroscopy. In the early to mid-1980s, Raman spectroscopy was considered to be a purely academic research tool with some industrial applications acknowledged by the academic community. Now, observing the composition of practitioners in the field of Raman spectroscopy it can be clearly seen that there are two groups. The first is composed almost exclusively of academics and scientists working in government laboratories, while the second comprises academics, industrial problem solvers, process and control engineers, and quality control technicians who started to use or reuse Raman spectroscopy as a tool from the late 1980s. Copyright © 2001 by Taylor & Franeis Group, LLC Several texts have appeared over the last few years, that have concentrated on either theory or instrumentation, FT-Raman, and dispersive spectrometers. When we first con- ceived this book, there was not a single reference source whose basis was (a) Can Raman spectroscopy be of use? and if it can, (b) what type of instrumentation will allow the user to obtain the most useful data? We have tried to identify as many areas of topical interest as possible, that will be appropriate to the second group of Raman practitioners described above. We have drawn together and blended industrial and academic spectroscopists to produce a comprehensive text for modern Raman spectroscopists. In the early chapters, the principles of Raman theory, instrumentation, selection of appropriate instrumentation, important instrumental measurement parameters, the use of data for the practitioner's benefit, and extracting the important analytical information from recorded data are introduced and discussed. These early chapters, where appropriate, use specific examples to illustrate the necessary concepts. ‘The later chapters provide extensive coverage of applications (semiconductors, carbon, hard-disk analysis, catalysts, glasses, pigments, gemmology, environmental science, in vivo biomedical studies, pharmaceutical studies, biodynamics, process control, archaeology, fo- rensics, polymer films and fibers) where Raman spectroscopy has proven to be an appro- priate tool is provided. In addition, fundamental reviews of gaseous, liquid, and solid phase Raman analysis are separate topics of discussion. Again, these chapters incorporate specific case-studies which should provide useful guidelines for the scientist planning to embark on the use of Raman spectroscopy as well as for the more experienced spectroscopist who plans to expand the use of Raman spectrometry into newer applications. The miniaturization of Raman spectrometers is now being addressed, and research- grade miniature Raman units will soon be available commercially, The drive to put a Raman spectrometer on Mars by 2005, providing arguably the most remote sensing Raman experiment ever undertaken, is an illustration of how the versatility of Raman spectroscopy is now being expanded. There are reports of suitable field instruments being developed of only several kilograms in mass, and the Raman spectrometers proposed for extraterrestrial planetary explorations are the smallest yet constructed (~1 kg or less). Despite the re- quirements for small size and rugged construction one thing is clear—there can be little or no sacrifice in spectral-quality information (resolution and sensitivity) provided by these miniaturized spectrometers. The already much expanded application portfolio of Raman spectroscopy is clearly ever-increasing. Jan R. Lewis Howell G. M. Edwards 2001 by Taylor & Francis Group, LLC Contents Preface Contributors 10 ‘Theory of Raman Scattering Laurence A. Nafie Evolution and Revolution of Raman Instrumentation— Application of Available Technologies to Spectroscopy and Microscopy Fran Adar Raman Spectrometry and Its Adaptation to the Industrial Environment Joseph B. Slater, James M. Tedesco, Ronald C. Fairchild, and lan R. Lewis Raman Microscopy: Confocal and Scanning Near-Field Kurt J. Baldwin, David N. Batchelder, and Simon Webster Raman Imaging Patrick J. Treado and Matthew P. Nelson The Quest for Accuracy in Raman Spectra Charles K. Mann and Thomas J. Vickers Chemometrics for Raman Spectroscopy Jeremy M. Shaver Raman Spectra of Gases Heinz W. Schrétter Raman Spectroscopy Applied to Crystals: Phenomena and Principles, Concepts and Conventions David C. Smith and Constantin Carabatos-Nédelec Raman Scattering of Glass Constantin Carabatos-Nédelec Copyright © 2001 by Taylor & Franeis Group, LLC ul 12 13 14 15 16 17 18 19 20 2 2 24 Raman Spectroscopic Applications to Gemmology Lore Kiefert, Henry A. Héinni, and Thomas Ostertag Raman Spectroscopy on II-VI-Semiconductor Nanostructures Bianca Schreder and Wolfgang Kiefer In Vivo Raman Spectroscopy Gerwin J. Puppels, Tom C. Bakker Schut, Peter J. Caspers, Rolf Wolthuis, M. van Aken, A. van der Laarse, Hajo A. Bruining, H. P. J. Buschman, Martin G. Shim, and Brian C. Wilson Some Pharmaceutical Applications of Raman Spectroscopy Adrian C. Williams Low-Frequency Raman Spectroscopy and Biomolecular Dynamics: A Comparison Between Different Low-Frequency Experimental Techniques. Collectivity of Vibrational Modes Ole Faurskov Nielsen Raman Spectroscopic Studies of Ion—Ion Interactions in Aqueous and Nonaqueous Electrolyte Solutions Jose M. Alfa Environmental Applications of Raman Spectroscopy to Aqueous Systems Ted L. Williams and Timothy W. Collette Raman and Surface Enhanced Resonance Raman Scattering: Applications in Forensic Science W. Ewen Smith, C. Rodger, Geoffrey Dent, and Peter C. White Application of Raman Spectroscopy to Organic Fibers and Films Stephen Michielsen Raman Spectroscopy of Catalysts Israel E. Wachs Applications of IR and Raman Spectroscopy to the Study of Medieval Pigments Fernando Rull Perez Raman Spectra of Quasi-Elemental Carbon James L. Lauer Process Raman Spectroscopy lan R. Lewis The Use of Raman Spectroscopy to Monitor the Quality of Carbon Overcoats in the Disk Drive Industry Andrew Whitley © 2001 by Taylor & Francis Group, LLC 25 Raman Spectroscopy in the Undergraduate Teaching Laboratory Michael D. Morris 26 Raman Spectroscopy in the Characterization of Archaeological Materials Howell G. M. Edwards Copyright © 2001 by Taylor & Franeis Group, LLC Contributors Fran Adar, Ph.D. Worldwide Raman Applications Manager, Raman Division, Horiba Group, Jobin Yvon Inc., Edison, New Jersey Jose M. Alfa, Ph.D., C. Chem., F.R.S.C. Professor, Department of Physical Chemistry, Universidad de Castilla-La Mancha, Ciudad Real, Spain Kurt J. Baldwin, Ph.D. Raman Consultant, Department of Physics and Astronomy, University of Leeds, Leeds, England David N. Batchelder, Ph.D. Professor of Physics, Department of Physics and Astronomy, University of Leeds, Leeds, England Hajo A. Bruining, M.D., Ph.D. Professor, Department of General Surgery, Erasmus University and University Hospital Rotterdam “Dijkzigt,” The Netherlands H. P. J. Buschman Laboratory for Intensive Care Research and Optical Spectroscopy, Department of General Surgery, Erasmus University and University Hospital Rotterdam “Dijkzigt,” Rotterdam, and Leiden University Medical Center, Leiden, The Netherlands Constantin Carabatos-Nédelec, Ph.D. Professor, Centre Lorrain d’optique et Eléctronique des Solides, University of Metz, Metz, France Peter J. Caspers, M.Sc. Department of General Surgery, Erasmus University and University Hospital Rotterdam “Dijkzigt,” Rotterdam, The Netherlands Timothy W. Collette, Ph.D. Research Chemist, National Exposure Research Laboratory, U.S. Environmental Protection Agency, Athens, Geo: G. Dent, Ph.D., C.Phys., Minst.P. Avecia Lid., Hexagon House, Manchester, England Howell G. M. Edwards, M.A., B.Se., D.Phil. C.Chem., FR.S.C. Professor of Molecular Spectroscopy, Department of Chemical and Forensic Sciences, University of Bradford, Bradford, West Yorkshire, England Copyright © 2001 by Taylor & Franeis Group, LLC Ronald C. Fairchild, M.S.E. (CICE) Director, Software Development, Kaiser Optical Systems, Inc., Ann Arbor, Michigan Henry A. Hanni, Pr. Dr. Director, SSEF Swiss Gemmological Institute, Basel, Switzerland Wolfgang Kiefer, Prof. Dr. Professor, Department of Physical Chemistry, University of Wiirzburg, Wiirzburg, Germany Lore Kiefert, Ph.D. Assistant Director, SSEF Swiss Gemmological Institute, Basel, Switzerland James L. Lauer, Ph.D. Emeritus Professor and Visiting Scholar, University of California, San Diego, San Diego, California Ian R. Lewis, Ph.D. Research Products Manager, Spectroscopy Products Group, Kaiser Optical Systems, Inc., Ann Arbor, Michigan Charles K. Mann, M.S., Ph.D. Professor, Department of Chemistry, Florida State University, Tallahassee, Florida Stephen Michielsen, Ph.D. Associate Professor, School of Textile and Fiber Engineering, Georgia Institute of Technology, Alanta, Georgia Michael D. Morris, Ph.D. Professor, Department of Chemistry, University of Michigan, Ann Arbor, Michigan Laurence A. Nafie, Ph.D. Distinguished Professor, Department of Chemistry, University, Syracuse, New York Syracuse Matthew P. Nelson, Ph.D. Principal Scientist, ChemIcon Inc., Pittsburgh, Pennsylvania Ole Faurskov Nielsen, M.Sc., D.Se. Professor, Department of Chemistry, University of Copenhagen, Copenhagen, Denmark ‘Thomas Ostertag, Dipl.-Min. Freelance Petrologist, Freiburg, Germany Fernando Rull Perez, Ph.D. Professor, Department of Crystallography and Minerology, University of Valladolid, Valladolid, Spain Gerwin J. Puppels, Ph.D. Associate Professor, Department of General Surgery, Erasmus University and University Hospital Rotterdam “Dijkzigt,” Rotterdam, ‘The Netherlands C. Rodger, Ph.D. Avecia Ltd., Hexagon House, Blackley, Manchester, England Bianca Schreder, Ph.D. Professor, Department of Physical Chemistry, University of Wirzburg, Wurzburg, Germany Heinz W. Schritter, Drrer.nat. Professor, Department of Physics, Ludwig- Maximilians-Universitat Miinchen, Munich, Germany Tom C. Bakker Schut, M.Se., Ph.D. Research Fellow, Department of General Surgery, Erasmus University and University Hospital Rotterdam “Dijkzigt,” Rotterdam, ‘The Netherlands Copyright © 2001 by Taylor & Francis Group, LLC Jeremy M. Shaver, Ph.D. Senior Scientist, Eigenvector Research, Inc., Manson, Washington Martin G. Shim, Ph.D. Research & Technology, Digital Security and Controls, Ltd., Concord, Canada Joseph B. Slater, B.S.E.E. Manager, New Product Development, Kaiser Optical Systems, Inc., Ann Arbor, Michigan David C. Smith, M.A., M.Se., Ph.D. Professor, Laboratory of Minerology, National Museum of Natural History, Paris, France W. Ewen Smith, D.Se., FRSE Professor of Inorganic Chemistry, Department of Pure and Applied Chemistry, University of Strathclyde, Glasgow, Scotland James M, Tedesco, M.S.E.E. Senior Staff Scientist, Kaiser Optical Systems, Inc., Ann Arbor, Michigan Patrick J. Treado, Ph.D. President, Chemicon Inc., Pittsburgh, Pennsylvania M. van Aken Laboratory for Intensive Care Research and Optical Spectroscopy, Department of General Surgery, Erasmus University and University Hospital Rotterdam “Dijkzigt,” Rotterdam, The Netherlands A. van der Laarse Laboratory for Intensive Care Research and Optical Spectroscopy, Department of General Surgery, Erasmus University and University Hospital Rotterdam “Dijkzigt,” Rotterdam, The Netherlands ‘Thomas J. Vickers, Ph.D. Professor, Department of Chemistry, Florida State University, Tallahassee, Florida Israel E, Wachs, Ph.D. Professor, Department of Chemical Engineering, Lehigh University, Bethlchem, Pennsylvania Simon Webster, Ph.D, Raman Consultant, Department of Physics and Astronomy, University of Leeds, Leeds, England Peter C. White, Ph.D., C.Chem., FR.S.C. Senior Lecturer, Forensic Science Unit, University of Strathclyde, Glasgow, Scotland Andrew Whitley, Ph.D. Director, Raman Spectroscopy Group, Jobin Yvon Inc., Edison, New Jersey Adrian C. Williams, B.Se., Ph.D., C.Chem, M.R.S.C. Reader in Pharmaceutical Technology, Drug Delivery Group, School of Pharmacy, University of Bradford, Bradford, West Yorkshire, England ‘Ted L. Williams, Ph.D. NRC Postdoctoral Fellow, National Exposure Research Laboratory, U.S. Environmental Protection Agency, Athens, Georgia Brian C. Wilson, Ph.D. Professor, Department of Medical Biophysics, Ontario Cancer Institute/Photonics Research Ontario, University of Toronto, Toronto, Ontario, Canada Rolf Wolthius, M.Sc. Department of General Surgery, Erasmus University and University Hospital Rotterdam “Dijkzigt,” Rotterdam, The Netherlands Copyright © 2001 by Taylor & Franeis Group, LLC 1 Theory of Raman Scattering Laurence A. Nafie Syracuse University, Syracuse, New York 1. INTRODUCTION Raman scattering is a fundamental form of molecular spectroscopy [1,2]. Together with infrared (IR) absorption, Raman scattering is used to obtain information about the structure and properties of molecules from their vibrational transitions [3-5]. The theory of Raman scattering is more complex that the theory of IR absorption, but there are a number of close parallels between the two theories [6-10]. As is well known, IR absorption arises from a direct resonance between the frequency of the IR radiation and the vibrational frequency of a particular normal mode of vibration. The property of the molecule involved in the resonant interaction is the change in the dipole moment of the molecule with respect to its vibrational motion. IR absorption is a one-photon event. The IR photon encounters the molecule, the photon disappears, and the molecule is elevated in vibrational energy by the energy of the photon at the frequency of vibrational resonance. By contrast, Raman scattering is a two-photon event. In this case, the property in- volved is the change in the polarizability of the molecule with respect to its vibrational motion. The interaction of the polarizability with the incoming radiation creates an induced dipole moment in the molecule, and the radiation emitted by this induced dipole moment contains the observed Raman scattering. The light scattered by the induced dipole of the molecule consists of both Rayleigh scattering and Raman scattering. Rayleigh scattering corresponds to the light scattered at the frequency of the incident radiation, whereas the Raman radiation is shifted in frequency, and hence energy, from the frequency of the incident radiation by the vibrational energy that is gained or lost in the molecule. The polarizability is a tensor with two Cartesian components; one is associated with the inci- dent photon and the other with the scattered photon. The two photons are connected by a single quantum mechanical process, a coherent event, that makes Rayleigh and Raman scattering different from the two one-photon events of absorption followed by emission. An energy-level diagram is given in Fig. 1, which illustrates IR absorption and Raman scattering. In both cases the initial state is the zeroth vibrational level of the ground Copyright © 2001 by Taylor & Franeis Group, LLC ho, go — 0 INFRARED RAMAN ABSORPTION SCATTERING Figure 1 Energy-level diagram for infrared absorption and stokes Raman Scattering for a vibra- tional transition from g0 to gl. The scattering photon energy, fie, is shifted from the incident laser radiation energy by the infrared vibrational energy, fiex,, gained by the molecule. electronic state (g0) and the final state is the first vibrational level of the ground electronic state (g1). IR absorption achieves this state change in one step, whereas Raman scattering requires two steps involving photon energies that are well above that of the IR photon or energy of the vibrational transition. Also indicated in Fig. 1 are the excited vibronic states for the molecule labeled ev for the vth vibrational level of the eth electronic state. If the molecule gains vibrational energy as shown, the scattering is called Stokes Raman, whereas if the molecule loses vibrational energy (by starting from an elevated vibrational level), the process is known as anti-Stokes Raman scattering. To varying degrees, Raman scattering is sensitive to all the excited electronic states of the moiecule. If the incident photon energy approaches the transition energy of an excited electronic state, typically the lowest allowed of such states, the Raman scattering, changes from normal Raman scattering to resonance Raman scattering. The general theory of Raman scattering takes into account this complexity, but there are two important limits that reduce the complexity. One is the far-from-resonance (FER) limit. This is a common situation for colorless samples that have no electronic states in close proximity to the incident photon energy. The other is the single-electronic-state (SES) limit of strong res- onance with a single electronic state. In this limit, the incident photon energy is very close to, or falls within, the absorption band of an excited electronic state of the molecule, and the resulting resonance Raman (RR) scattering is dominated by the properties of thi resonant electronic state. In the past 25 years, a new area of Raman spectroscopy has evolved that is also a new form of molecular optical activity [9,11—13]. Called Raman optical activity (ROA), this form of Raman scattering applies only to chiral molecules, molecules whose mirror- image pairs are nonsuperposable. ROA is broadly defined as the difference in Raman scattering for right versus left circularly polarized (CP) radiation, where the change in the CP state of the light is effected for either the incident radiation, the scattered radiation, or both, either in phase or out of phase. The theory of ROA is more complex than the theory Copyri 2001 by Taylor & Francis Group, LLC of ordinary Raman scattering because it involves additional terms in the expansion of the polarizability beyond the dipole approximation. The new optical activity tensors involve either the magnetic dipole moment or the electric quadrupole moment, In this chapter, we will provide a very basic, unified treatment of the theory of Raman scattering and ROA. More extensive descriptions of theory of Raman and ROA scattering can be found elsewhere [6-10,12,14,15] il. THE RAMAN TENSOR One of the essential properties associated with the Raman scattering of light by molecules is the polarization state of the light [9]. Changes in the polarization state affect the nature and information content of the scattered light. The intensity of light scattering for any experiment can be expressed in terms of the polarizability, &,,, and the polarization vectors for the incident and scattered radiation, é!, and é4, respectively, and is given by We", 90K 22 a.p2l") wo In this equation, the angular brackets designate an average over all angles of orientation of the molecule to the laboratory frame of reference. This is needed for liquid, solution, or gaseous samples where there is no unique orientation of molecular axes relative to the laboratory axes. The polarization vectors have one Greek subscript and the scattering tensor has two. For repeated Greek subscripts, summation over the Cartesian directions x, y, and z is implied. Hence, Eq. (1) has, in general, nine terms (one for each component Pair xx, xy. 42, JM, Ys Yes Lk ZY, and zz) within the vertical brackets, and these brackets with superscript 2 designate the absolute square of the complex quantities within the brackets, The tilde above a quantity, such as a polarization vector or a scattering tensor, indicates that this quantity can be complex. The asterisk superscript for the polarization vector of the scattered light designates complex conjugation. The constant K is given by L fwtwk 90. ( 4aR ) 2 where w is the angular frequency of the scattered light, to is the magnetic permeability, E° jg the electric field strength of the incident laser radiation, and R is the distance from the scattering origin to the detector. The Raman polarizability tensor is given by [12,14] By (qeeualale ' (nb lf) @ OF Wp, ~ @y + it, m+ Oy + iD, where fi is Planck’s constant divided by 277, and the summation is over all excited elec- tronic states, j, of the molecule. The states n and m differ by a vibrational quantum of energy. The denominators contain frequency terms, and @, is the angular frequency dif- ference between the states j and n. The terms iT’, are imaginary terms proportional to the width of the vibronic state j and, hence, inversely proportional to its lifetime. The first of the two terms in Eq. (3) is called the resonance term because the frequency difference between the jn transition frequency and the laser frequency vanishes at the resonance condition. The quantities in angular brackets are quantum mechanical matrix elements with electric dipole moment operators fi,, given by Copyright © 2001 by Taylor & Franeis Group, LLC =D) etn 4) which is simply the summation over the charge and position in the ath direction of all particles, k, in the molecule, electrons, and nuclei. For Raman scattering, only the electrons of the molecule need to be included in the sum because the nuclei make no contribution to the polarizability. The matrix element in Eq. (3) involving the operator ji, describes the interaction of the molecule with the incident radiation, and the matrix elements with the operator /i,, describes the interaction of the molecule with the scattered radiation, The matrix element products in each term can be read from right to left in a time-ordered sense; hence, the resonance term describes the molecule interacting first with a laser photon and subse- quently creating a scattered photon, whereas the nonresonance terms reverses the natural order of those two events. Ill. RAMAN INTENSITY INVARIANTS The general theory of Raman scattering embraces all possible polarization experiments, scattering geometries, and degrees of resonance Raman-intensity enhancement. Raman intensity is proportional to the square of a tensor quantity, as expressed in Eq. (1). For solid samples whose crystal axes are aligned with laboratory axes, this equation is suffi- cient for practical use. However, for most applications by chemists, the sample is in a state where the molecules are oriented randomly with respect to the laboratory axes, such as a gas, a liquid, a solution or a powder. In these cases, an orientational averaging must, be carried out for the sample molecules. The details of this averaging are complex and can be found elsewhere [9,16]. After the averaging has been carried out, there are special linear combinations of polarizability tensors whose values are independent of the orien- tation of the molecule relative to laboratory frame. Such combinations are called invariants. These are important quantities in Raman theory because polarizability tensor expressions can be written down, or calculated, for the molecule independent of its orientation. Other factors then take care of changes in the laboratory conditions, such as scattering geometry and polarization states for the incident or scattered light. It turns out that all Raman intensities from samples of randomly oriented molecules can be expressed in terms of only three invariants, called the isotropic invariant, the symmetric anisotropy, and the antisymmetric anisotropy, and these are given by [15,16] J Ref(nn)'(Ggg)*] 6) B,(a)* = $Re{3(Gag) ap)" — ()"(p)*1 © Ba)? = Re[34,5)"(Gag)*1 a where the symmetric and antisymmetric forms of the polarizability tensors are given by (Gp) = (Gap) + (Gp) @) (Gp) = [UGup) ~ G01 @ As before, a repeated Greek subscript in a tensor or tensor product indicates summation 2001 by Taylor & Francis Group, LLC over the Cartesian directions x, y, and z, Thus, in Eqs. (5)—(7), there are nine terms to consider to write out the full expression. Using these invariants, a general theory of Raman scattering can be constructed that describes all Raman experiments and depolarization ratios. IV. GENERAL THEORY OF RAMAN OBSERVABLES Any Raman scattering experiment, for samples of randomly oriented molecules, can be expressed in terms of the three invariants given in Eqs. (5)—(7). We provide here expres- sions for a few of the more common cases involving right-angle scattering and backscat- tering. For laser radiation incident along the laboratory positive Z axis, three basic seat tering geometries can be defined: forward scattering (0°) along the positive Z axis, right-angle scattering (90°) along the ¥ axis, and backscattering (180°) along the negative Z axis. In the case of right-angle scattering, there are two important scattering experiments. ‘They are the depolarized case where the incident light is X-polarized while the scattered light is Z-polarized and polarized scattering where both the incident and scattered beams are linearly polarized in the X direction. These are given by Eqs. (10) and (11); the superscript and subscript on the intensity symbol, /, indicate the direction of the polari- zation of the incident and scattered light, respectively, and the scattering angle is indicated in parentheses: T(90°) = 2K[3B.(a)" + 5B,(@)"] (10) Ti(90°) = 2K [450° + 48,(a)"] ap The ratio of these two intensities is the depolarization ratio, namely 3B.(a)" + 5B,(a)" 45a? + 4B,(a)° The depolarization ratio varies from zero for Raman scattering from totally symmetric modes where only the isotropic invariant is nonzero, through 3/4 for nontotally symmetric modes in the i nonzero, and on to © for modes where only the antisymmetric anisotropic invariant is nonzero. It can be shown that the antisymmetric anisotropic invariant can be nonzero only for resonance Raman scattering. In general, it is not possible to isolate all three Raman tensor invariants from the two right-angle scattering measurements, and a third linearly independent measurement is needed. This can only be achieved using a different scattering geometry and circularly polarized radiation. Two convenient options are backscattering using purely circularly po- larized radiation (right or left) where either the sense of the circularly polarized radiation of the incident and scattered beams is the same (RR or LL) or the opposite (RL ot LR). For the case of incident right circularly polarized radiation, we have (2) Pp (180°) = 2K [6B,(a)"] 3) TN(180°) = 2K [4502 + B(a)* + 58,0") (4) The ratio of these two intensities is called the reversal ratio and is given by 6Ba)” * Ba + Bla + 5Blay a The Raman intensity in Eq. (13) corresponds to corotating circularly polarized backscat- Copyright © 2001 by Taylor & Franeis Group, LLC tering and is a purely depolarized experiment that includes no contribution from either the isotropic invariant or the antisymmetric anisotropic invariant. The scattering in Eq. (14) corresponds to contrarotating circularly polarized radiation for the incident and scattered beams and is a form of polarized Raman scattering. V. FAR-FROM-RESONANCE LIMIT The theory of Raman scattering simplifies dramatically in the far-from-resonance (FFR) limit, where the exciting laser radiation is far from the lowest allowed excited electronic state of the molecule [15]. In this limit, the interaction of the light with the molecule is approximately the same for both the incident and the scattered radiation. There, the Raman tensor becomes symmetric and this symmetry reduces the number of Raman invariants from three to two, the isotropic and (symmetric) anisotropic invariants. The antisymmetric anisotropic invariant vanishes, by its antisymmetric definition in Eqs. (7) and (9), and we have Bay = 0 (16) Baa = Play ay The equations for the two nonzero Raman invariants are C= ancttgy (18) BAY? = } Be, 564p — Haan) (a9) where the FFR polarizability is given simply by pe wt Relend alMAaaln)] (20) Here, only the real part (Re) of the matrix element product is needed. This expression can be compared to the corresponding general expression for the polarizability given in Eq. (3). Using these invariants, we can write intensity expressions for Raman scattering that cover all possible polarizations and scattering geometries in the FFR approximation. The Raman-intensity expressions for right-angle scattering and the depolarization ratio in the FER limit are given by 15(90°) = 2K[3B(a)'] Ql) 15(90°) = 2K[45a? + 4B(a)"] (22) po 3 Play" (23) ~ 45a + 4B(ay The depolarization ratio given in Eq. (23) is the more familiar form where the allowed values range from zero for a spherically polarized band to 3/4 for any non-totally s metric vibration, where the isotropic invariant is zero for the latter. The corresponding expression for corotating and contrarotating backscattering in the FFR limit are given by Ti(180°) = 2K [6B(a)"] 24) (180°) = 2K[45a* + PCa) (25) m- Copyright © 2001 by Taylor & Francis Group, LLC 45a’ + Bla) 28) Here, the reversal ratio varies from 0 to 6 for the same limits as the depolarization ratio in Eq. (23). The above intensity expressions may be combined for other experimental setups in right-angle scattering and backscattering. For example, for incident linearly po- larized radiation in right-angle scattering and unpolarized scattering radiation (all scattered radiation measured, both X- and Z-polarized), one hi £590") = 2K[45e? + 7B(a)"] (27) Similarly, for backscattering, the same intensity if obtained by summing Eqs. (24) and (25). Here, for unpolarized light collection, the scattered intensity is independent of the polarization state incident radiation: Ti80°) = F180") = 2K[45a° + 7B(a)"] (28) Note that in the FFR limit, the value of both Raman intensity invariants may be obtained by using only two different polarization measurements and there is no need to change the scattering geometry, as there is the general case in order to isolate three Raman invariants. We also note that the distribution of the a is different between the polarized and depolarized forms of scatter ig and backscattering geometries even though the (otal scattered intensity is the same. The pure backscattered anisotropic scattering is wwice the value for the corotating backscattering compared to the perpendicular right-angle scattering [compare Eqs. (21) and (24)]. For this reason, there may be advantages to setting up backscattering compared to right-angle scattering when one wants to emphasize the difference between the scattering from the isotropic invariant and that of the anisotropic invariant. VI. SINGLE-ELECTRONIC-STATE RESONANCE LIMIT When the frequency of the incident laser radiation in a Raman scattering experiment is in resonance with a single electronic state (SES), it is well known that strong enhancement of the Raman scattering occurs [14]. This is because the denominator of the resonant term in the polarizability expression, Eq. (3), approaches zero and the value of the polarizability can increase by several orders of magnitude. The resonance condition brings simplifying conditions to the theory of Raman scattering, and the phenomenon itself becomes known as resonance Raman (RR) scattering. Under conditions of strong resonance, the nonreso- nant term {second term in Eq. (3)] can be dropped and the contributions of all other electronic states can be omitted as too small to consider. The Raman polarizability in Bq (3) for a fundamental vibrational transition in the ground electronic state, 0 to gl, becomes - _ Ly (ellalevXevl pl 30) Cageue =F Dg oe tay ” If the transition moment of the resonant electronic state is taken to lie in the z direction, the general set of three Raman invariants reduces effectively to only one Raman invariant as Copyright © 2001 by Taylor & Franeis Group, LLC 5M@Dar.col® (30) Baa)? In addition, it can be shown that this Raman invariant is proportional to the square of the (Ger.cal” BD electronic absorption strength for the resonant electronic state through the relationship < 1) a0 4 [Gdeical (5) [Ge |*U (eo) (32) where (ji), is the electric dipole transition moment between the ground and the resonant electronic state, respectively. From this relationship, it is clear that there is an underlying connection between RR scattering in the SES limit and the electronic absorption spectrum, of the resonant electronic state. If one varies the excitation frequency of the incident laser beam and measures the relative intensity of the RR scattering for any particular band, one obtains a spectrum called the excitation profile of the band, which carries the resonance dependence of that band, The form of the frequency dependence of the excitation profile can be seen by taking the absolute square of the denominator of Eq. (29). Usually, the resonant vibronic states that are most important for a given vibration are the e0 and el vibronic bands. This assumes that the excited-state potential surface is nearly the same as that of the ground state except for a shift in the equilibrium nuclear position. If more than one electronic state participates in the resonance enhancement of bands, either by direct resonance or indirectly through vibronic coupling, excitation profiles can reveal informa- tion about the origin and symmetry of individual RR bands. Vil. RAMAN OPTICAL ACTIVITY A relatively new form of Raman scattering combines ordinary Raman scattering with natural optical activity. Called Raman optical activity (ROA), this form of Raman tering arises as the difference of the Raman scattered intensity for left versus right cir- cularly polarized radiation, either in the incident light, the Raman scattered light, or both [9,11,12]. In order to describe ROA theoretically, one needs to consider the next higher interactions of light with matter. One is the magnetic dipole moment whose operator is closely related to the electric dipole moment operator in Eq. (4) and is given by & = Bn, Set "0Poy (33) This expression contains the vector cross-product, written in Cartesian notation, of the position of the electron rand its momentum p. Here €,9, is the antisymmetric tensor which equals —1 for an odd permutations of the Cartesian axes x, y, and z, and +1 for an even permutation. Similarly, one must consider the electric quadrupole moment operator given in Cartesian tensor notation by ; ee , bug = >) F Grrarap ~ 77806) (G4) where 8,, is the Kronecker delta function which is equal to +1 if the Cartesian subscripts are equal and zero if they differ. These operators can be substituted into the general expression for the polarizability given in Eq. (3) to yield the magnetic dipole and the electric quadrupole optical activity tensors. In the FFR limit, simpler expressions for these 2001 by Taylor & Francis Group, LLC tensors are obtained that are analogous to the FFR polarizability tensor given in Eq. (20) and are given by =2 Gly ea limba h ln) 35) Rel(n| a |/ XJ] Bool n)I (36) om “a where Im and Re represent the imaginary and real parts, respectively, of the expressions in brackets that follow. From these tensors arise three ROA invariants that are analogous to the two FFR Raman invariants given in Eqs. (18) and (19). The three ROA invariants are the magnetic dipole isotropic invariant and the magnetic dipole and electric quadrupole anisotropic invariants, given respectively by AG! =F n,Gin (7) BG'Y = {B0gGi9 ~ GaaGbe) (38) BAY = 5 dust aysAyas 9) The two most important experimental setups for the measurement of ROA are backscat- tering (180°) incident circular polarization (ICP) ROA and backscattering in-phase dual circular polarization (DCP,) ROA. The ROA and parent Raman intensities written in terms of ROA and Raman invariants in the FFR limit are given by T8(480°) — 1.180") = *« f12pi@’y + 4p(Ay’] (40) TS(180°) + (180°) = 4K [4507 + 7B(a)'] (41) respectively, for backscattering ICP ROA where the subscript u represents unpolarized scattered light and the superscript represents the polarization state of the incident laser radiation. For backscattering DCP, Raman and ROA intensity, the expressions are Ts(180°) — 17180") = * (12B(G'Y + 4B(ay'] (42) 18(180°) + 1E(180°) = 4K [6B(a)"] (43) ‘The ROA intensities are the same for these two forms of ROA, but the corresponding Raman intensities differ. This is because ICP ROA does not discriminate the polarization state of the scattered light and additional Raman scattering is measured that carries no ROA in the FFR limit, namely the out-of-phase DCP,, Raman and ROA intensities. Back- scattering ICP, and DCP, ROA are the most efficient ways to measure ROA spectra that can be devised, and in the last decade, nearly all new measurements of ROA have been carried out using one or the other of these two setups. VII. AB INITIO CALCULATIONS OF RAMAN INTENSITIES: Raman jntensities can now be calculated for any molecule for which the equilibrium structure and vibrational force field can be determined. This has been achieved [17] by Copyright © 2001 by Taylor & Franeis Group, LLC starting with expressions for the polarizability in the zero-frequency limit of the FFR approximation as 2S Relin| fi.) / Mil Hp!n)) == a (4d) Aug = 5 2 on (aay where the distinction between the initial and final states is not needed, neither is the frequency of the incident laser radiation in the frequency denominator. The Raman inten- sities can be obtained from these tensors by calculating their variation with the normal coordinates of vibrational motion. The method by which these tensors have been calculated s the electric field perturbation approach. The summation over all the excited states j can be avoided by substituting field perturbed wavefunctions in first-order perturbation theory for their nonperturbed counterparts as 2 >) Relin|i.ln'(Ep)1 (45) where E, is the ath Cartesian component of the electric field, and the prime on the wave function indicates the first derivative with respect to the field. Using ab initio quantum mechanical methods, excellent agreement between theory and experiment has been achieved. A particularly successful approach is the use of density functional theory (DFT) with the hybrid density functionals B3LYP and BPW94 and basis functions at the level of 6-31G* or higher. Once the Raman tensors are calculated, invariants can be constructed using Eqs. (18) and (19), and then assembled into the combinations appropriate for the experimental setup desired. Extensions of these calculations have also been carried out for ROA, where good agreement with experiment has been achieved. REFERENCES G Herzberg. Infrared and Raman Spectra of Polyatomic Molecules. New York: Van Nostrand, 1945, EB Wilson, JC Decius, et al. Molecular Vibrations. New York: McGraw-Hill, 1955. JG Grasselli, BJ Bulkin, eds. Analytical Raman Spectroscopy. New York, Wiley, 1991 JR Ferraro, K Nakamoto. Introductory Raman Spectroscopy. New York: Academic Press, 1994. B Schrader, ed. Infrared and Raman Spectroscopy. New York: VCH Publishers, 1995. MC Tobin, Laser Raman Scattering. New York: Wiley, 1971 JA Koningstein, Introduction to the Theory of the Raman Effect. Dordrecht: Reidel, 1972. DA Long. Raman Spectroscopy. New York: McGraw-Hill, 1977, LD Barron. Molecular Light Scattering and Optical Activity. Cambridge: Cambridge University Press, 1982. 10. WB Person, G Zerbi, X Qu, TB Freedman, eds. Vibrational Intensities of Infrared and Raman Spectroscopy. Amsterdam, Elsevier, 1982. 11, LA Nafie, GS Yu, Comparison of IR and Raman forms of vibrational optical activity. Faraday Discuss 99:13~34, 1994. 12. LA Nafie. Theory of resonance Raman optical activity: The single-electronic state limit. Chem Phys 205:309-322, 1996. 13, LA Nafie. Vibrational optical activity. Appl Spectrosc 50(5):14A-26A, 1996. 14. PM Champion, AC Albrecht. Annu Rev Phys Chem 33:353, 1982. 15. LA Nafie, D Che. Theory and measurement of Raman optical activity. In: M Evans, $ Kielich, eds. Modern Nonlinear Optics, Part 3. New York: Wiley, 1994, vol 85, pp 105-149. 16, L Hecht, LA Nafie. Theory of natural Raman optical activity I. Complete circular polarization formalism, Mol Phys 72:441-469, 1991 17, PL Polavarapu. Ab initio vibrational Raman and Raman optical activity spectra. J Phys Chem 94: 8106-8112, 1990. Serva S 2001 by Taylor & Francis Group, LLC 2 Evolution and Revolution of Raman Instrumentation— Application of Available Technologies to Spectroscopy and Microscopy Fran Adar Jobin Yvon Inc., Edison, New Jersey 1. DISCOVERY OF THE RAMAN EFFECT Chandrasekhara Venkata Raman first reported in 1928 the light-scattering phenomenon that has come to be known as the Raman effect [1]. This first publication represented the culmination of his research performed during the entire decade of the 1920s [2]. In fact, there had been earlier descriptions of phenomena now recognized to be examples of the Raman effect. In 1878, Lommel [3] described characteristics of fluorescence, which were dependent on the identity of the sample and the frequency of the exciting radiation, On the theoretical side, Smekal [4], Kramers and Heisenberg [5], Schroedinger [6], and Dirac (7| applied quantum mechanics to molecules during this period when quantum mechanics was under development and they predicted the Raman effect. Raman’s interests in physics until this time had covered the fields of acoustics as well as optics, and his research efforts in light scattering were initiated by his observations of the color of the sky and Mediterranean Sea during his travels to and from the 1921 Congress of Universities of the British Empire held in Oxford. Rayleigh had already attributed the sky's color to scattering of sunlight by air molecules. Raman, however, showed experimentally that the color of the sea results from intrinsic light scattering by water molecules rather than reflection of the light from the sky, as conjectured by Rayleigh. Already in 1922, Raman published an article with these observations and ideas [8]. In a 1922 memoir, entitled “The Molecular Diffraction of Light,” he laid out the physical concepts and issues regarding molecular scattering and diffraction, including the quantum theory of light, together with a description of measurements already performed with his group, and plans for future experiments. In the years between 1923 and 1925, Copyright © 2001 by Taylor & Franeis Group, LLC they studied the scattering of light by water and other liquids, confirming a “feeble fluo- rescence” in all cases, but attempts at spectroscopic analysis had failed. By the end of 1927, Raman had come to the conclusion that the light that he was observing was an optical analog to Compton scattering of electrons [9]. Following this, insight, he and Krishnan, his student, studied a number of liquids, organic vapors and gaseous CO; and N;O, as well as some crystals and amorphous solids. Their observations were described in detail in an address delivered to the South Indian Science Association in Bangalore on March 16, 1928 [2]. In a second 1928 publication in Nature, he reported the details of their experiments showing that “the spectrum of the new radiation consist(s) mainly of a narrow range of wavelengths clearly separated from the incident spectrum by a dark space” [10]. Using a mercury arc lamp, they were able to determine that any exciting line was accompanied by weaker lines satisfying the now famous equation With this equation, Raman recognized that there was energy exchange between the incident photon and internal excitations of the scattering medium, corresponding to the partial exchange of energy into atomic vibrations of the molecules. Raman’s earliest spectra were recorded with a small prism spectroscope equipped with a photographic plate. The spectra were excited with a mercury arc lamp, using a filter to select the Hg line of interest, and the samples were contained in a large spherical flask. ‘A photograph of his spectrograph appears in the January 18, 1999 issue of C&ENews. During this period, other laboratories around the world were also studying light scat- tering. Jayaraman and Ramdas [2] cite the younger Lord Rayleigh, Wood, Cabannes, and Landsberg and Mandelstam. More details about the contributions from these laboratories can be seen in the introductory chapter of Bernhard Schrader’s text Infrared and Raman Spectroscopy [11]. Schrader notes that Raman cabled his original manuscript to Nature on February 16, 1928. According to Schrader, Rocard immediately assigned Raman’s obser- vations to optical modulations of oscillating molecules. Landsberg and Mandelstam, work- ing in the U.S.S.R. Academy of Sciences in Moscow reported the effect in crystalline quartz, and calcite almost simultaneously [12]. Due to controversy, the Russian literature refers to the inelastic light-scattering phenomenon as “combinatorial scattering” [13] Bernhard Schrader also reviewed the earliest observations of vibrational spectra, de- scribing molecular absorption in the infrared part of the spectrum as well as Raman scat- tering. Even before the turn of the century, at least four workers had determined that absorbance spectra were molecule-specific. In 1905, Coblentz [14] published an atlas of absorption spectra of 120 organic compounds using equipment he built himself. Coblentz’s text even included a description of the role played by molecular functional groups. How- ever, it is generally acknowledged that vibrational spectroscopy was not actively applied to analytical chemistry until after the Second World War when both Raman and infrared GR) instrumentation began to evolve into commercial, easier-to-use tools. Until 1961, the Raman effect was applied to the study of the vibrational spectra of gases, liquids, and solids, and the rotational spectra of gases. The study of these excitations with visible light was much easier than through direct absorption in the mid-infrared and far-infrared part of the spectrum, However, it was early recognized that Raman scattering and infrared absorption were quite complementary because of the differences in two- photon versus one-photon coupling mechanisms between the molecules and the photons. In fact, in some cases, the differences in selection rules came to play a role in identifying Copyright © 2001 by Taylor & Francis Group, LLC molecular structure where several structures had been proposed. The availability of lasers in the early to mid-1960s further stimulated the application of the technique to many solid- state phenomena as well. This is where our story of instrumentation development will start, Note: This historical account of the discovery and early use of the Raman effect is based on the article cited written by Jayaraman [2], an associate of Raman (between 1949 and 1960) and Anant Ramdas, his last student (1950-1956), on the introductory chapter to Schrader’s book [LI], and articles by Brandtmuller and Keifer [15] and Long [16]. As the reader will see, the evolution in Raman instrumentation did not follow a linear path. The instrument evolution demonstrates significant interactions among the roles played by the sample properties, the excitation wavelength, evolution of the detectors, monochromator design driven to match the detectors, implementation of FT-Raman (Fou- rier transform Raman), and the surprising impact of the Raman microprobe. As each development is described, there will be small diversions included in square brackets, al- most like a point-counterpoint musical exposition, that explain the interplay of these developments. The selection of topics to be included was motivated to expose the evolution of commercial instrumentation for applications in materials science and analytical spec- troscopy. Traditionally, there have been numerous other Raman topics that have been quite interesting but not included here because of our selected focus. The reason for the slow development of sensitive instrumentation for Raman analysis, is the feebleness of the Raman signal itself. As the reader will see, the evolution of equipment that enable the routine recording of spectra in seconds-to-minutes versus hours- to-days took about a half of a century. The Raman effect is a weak process, at least in part because it is a second-order process (meaning that two-photon interactions are in- volved). For this reason, the introduction of the laser provided a major breakthrough. The laser is not only orders of magnitude more intense than any incoherent source, but the light comes out as a narrow, collimated beam, which is easy fo “plumb” quite efficiently into sampling optics, However, the intense laser source was not the only important development. Because the Raman effect is weak, there is often a very intense contribution to the scattered beam at the laser wavelength, accompanying the Raman signal. This signal is often referred to as the Rayleigh signal, even though, strictly speaking, the Rayleigh signal is an intrinsic signal, not a signal dependent on the optical properties of the sample. Until the relatively recent availability of holographic notch filters, multiple-stage monochromators or spectro- graphs were necessary in order to separate the signal of interest from the intense laser Rayleigh”) signal. The optical systems that were used were tailored to the detectors. Section II discusses the characteristics of “scanning” instruments in which each wave~ length of the Raman spectrum is presented sequentially to a “single-channel” detector through the exit slit of a double (or triple) monochromator, Section III discusses the characteristics of multichannel detectors and triple spectrographs. Section IV discusses the importance of the holographic notch filters in enabling the design of single spectrographs. Section V covers interferometric instruments (Fourier transform Raman) that eventually also used holographic filters to eliminate the laser wavelength. Section VI discusses the microprobe. Section VII covers briefly remote sensing and instruments designed specifi- cally for monitoring manufacturing processes. In Section VIII is a summary of the con- siderations involved in the choice of laser excitation wavelength, and in Section IX is a chapter summary and indication of how analytical instrumentation may evolve in the near future. Copyright © 2001 by Taylor & Franeis Group, LLC Il. LASERS, DOUBLE MONOCHROMATORS, AND PHOTON-COUNTING PHOTOMULTIPLIER TUBES A. Background The challenge to the development of “practical” Raman instruments was to enable the collection of signals from highly scattering solids as well as liquids and gases, in reason- able times. The instruments that were used most widely between the mid-1960s and the mid-1980s were double scanning monochromators with laser sources and low-noise pho- tomultipliers (PMTs). Immediately following the invention of optical lasers in the early 1960s, all instru- ments used lasers as the light source. Acquisition of spectra from samples with arbitrarily high “Rayleigh” radiation required some spectrometer developments as well as imple- mentation of low-noise detectors. Curiously, the physics on which both laser and photo multiplier operation are based is derived from work of Albert Einstein. Photomultiplicrs operate by using the photoelectric effect for which Einstein received the Nobel Prize for work done in 1905. Laser action was a demonstration of stimulated emission that was predicted by Einstein in 1917. A discussion of laser action will not be included in this chapter, but the reader can refer to a text on optoelectronics such as Ref. 17. ‘One of the curiousities of Raman instrumentation is the manner in which the detection method has evolved and, in fact, come full circle (from small spectrographs with photo- graphic plates, to scanning monochromators with sensitive single-channel photomultipliers, to spectrographs with sensitive, low-noise multichannel detectors). Raman’s instrument was a small spectrograph with a photographic plate. Similar instruments used in the early years typically required more than 12 hr to acquire spectra [18] B. First Laser-Excited Spectra Until the introduction of visible lasers, Raman spectra were generated with low-pressure, water-cooled mercury arc lamps. As early as 1961, Townes recognized the usefulness of using a laser to excite Raman spectra [19]. Porto and Wood almost immediately applied a pulsed ruby laser to excite the Raman spectra of liquids [20] in 1962. The continuous- wave (CW) HeNe laser was implemented in at least two laboratories in the next 2 years [21,22]. These authors discuss the advantages of the CW HeNe over the pulsed ruby laser. Weber and Porto were the first to use a HeNe laser as a source for Raman spectra of gases. In order to compensate for the low gas volume that would be illuminated, they designed an intracavity scheme to take advantage of the much higher power density in the laser resonator. They were able to measure pure rotational scattering (Stokes and anti- Stokes) close to the laser line on an echelle spectrograph equipped with a photographic plate. C. Monochromators and Stray Light During the period between the 1960s and the 1980s, the monochromators were “scanned” in order to take advantage of the low-noise photomultipliers (PMTs) developed by astron- omers. For stray light rejection, multistage monochromators were developed. The “stray light” passed to the detector was effectively reduced by intermediate slits between the monochromator stages (23]. Both two-stage (Cary, Jarrel-Ash, Spex, Lirinord/Jobin Yvon/ ISA) and three-stage (Coderg) systems were available, but most systems delivered had only two stages. Depending on the optical design, the dispersion of the second mono- Copyright © 2001 by Taylor & Francis Group, LLC chromator could “add” to the dispersion of the first or “subtract” (i.e., cancel) it. A discussion of the effect of the monochromator design on stray light showed that, everything else being equal, the stray-light rejection would be better in a double subtractive rather than a double additive monochromator. This follows because the slit areas over which the “stray light” is integrated are smaller in double subtractive systems. To further reduce the “stray light,” holographically recorded gratings were imple- mented in Raman instruments in 1973 [23,24]. Previous to this development, gratings that were produced with a diamond tool scribing lines on the grating surface were susceptible to the introduction of periodic defects due to errors in the lead screw driving the cutting and/or thermal drifts in the room in which the cutting was being done. The holographic process provided a method that inherently avoided these errors. In addition, the surface structure was more “perfect.” The final result of these differences produced gratings with- out “ghosts” (light diffracted by the periodic errors) and with stray-light levels reduced by orders of magnitude. Because of the feebleness of the Raman signal, photon-counting PMTs were imple- mented in the systems developed in the 1960s. The attractiveness of PMTs arose from their high sensitivity, very wide dynamic range, and linearity of response. The history and characteristics of these devices is reviewed in a technical book published by Burle/RCA (25]. To quote this monograph [26, p. 3], “The basic reason for the superiority of the photomultiplier is the secondary-emission amplification that makes it possible for the tube to approach ‘ideal’ device performance limited only by the statistics of photoemission.” That means that with amplifications between 10° and 10°, signals are not swamped by auxiliary electronic equipment noise; in addition, extremely fast response times (10™'" sec, for following fast kinetic processes) can be achieved. Astronomers were attracted to the devices because of the effective quantum efficiencies that were seen to be 10 times that of photographic film. The linearity of the response and the wide range of gain made these detectors particularly useful for recording line spectra. Including them in Raman systems followed naturally. D. Operating Characteristics of a PMT Following absorption of a photon, electrons are emitted from a metal with kinetic energy equal to the difference between the photon energy and the metal’s work function, Because it is a quantum process, the photoelectric current is proportional to the intensity of the radiation. The peak of the wavelength response of the alkali metals depends in a monotonic fashion on the position of the metals in the periodic table. Li peaks at about 390 nm, and Cs peaks at about 560 nm. ‘There are some limitations to the effectiveness of metal emitters that were improved by the introduction of semiconductor photocathodes. The disadvantages of metal emitters include the following: + High reflective losses electron scattering due to large numbers of free + Small escape depth (few nanometers) also due to large electron—electron scattering + Large work function (>3 eV typically) allowing detection of photons only in the blue and near ultraviolet (UV). In semiconductors, photons can only be absorbed (detected) if their energy is greater than the bandgap (energy difference between the valence and conduction bands) plus the Copyright © 2001 by Taylor & Franeis Group, LLC electron affinity (the energy required to release the electron from the surface into the vacuum of the tube). In semiconductors, light can be detected throughout the visible by using a variety of semiconductors whose bandgaps fall in the visible to near-infrared (NIR) part of the spectrum (silicon and gallium arsenide, for examples). The fact that electron scattering is considerably lower than in metals because of the much lower number of carriers means that the escape depths can be tens of nanometers, much larger than that of metals. In addition, by treating the surface of the semiconductor, it became possible to make the electron affinity less than zero, which increased the electron escape depth by two orders of magnitude. In summary, whereas the earliest Raman measurements were recorded on spectro- graphs with low-pressure mercury sources and photographic plates (the earliest “multi- channel” detector), scanning multistage monochromators with lasers and PMTs became the standard instrument during the 1960s and 1970s. These systems provided high stray- light performance, allowing examination of all types of samples, including highly “s tering” solids. Raman bands at shifts as low as 5-10 cm_' from the exciting line were observed, Il. MULTICHANNEL DETECTORS AND TRIPLE SPECTROGRAPHS With the development of low-noise multichannel detectors in the 1980s [26}, the instru- mentation used to measure Raman signals went through another revolution, In order to take best advantage of the “real estate” of the multichannel detectors, the spectrographs were redesigned. This meant that instead of using long-focal-length double monochro- mators which did a superb job of suppressing the stray light resulting from the high Rayleigh signals, one really needed much shorter focal lengths in order to display reason- able amounts of spectra (with acceptable spectral resolution) on the 0.5-in. or 1-in. arrays. ‘The instruments introduced during the 1980s were “triple spectrographs,” in which the first two stages were used in subtractive dispersion [27]. The intermediate slit was set to select the spectral range viewed by the spectrograph and detector, with the central wave- length carefully selected to assure that the laser wavelength did not pass into the spectro- graph stage. A. Triple Spectrographs The properties of this type of triple system have been carefully described [27]. The size of the systems (i.e., the focal lengths) were determined by the desired dispersion on the array. For general chemistry applications, 1 cm™'/detector pixel seems to be the range of what the systems need to deliver. That provides an approximate full width at half-maxi- mum (FWHM) of 3 om”, For use with lasers producing wavelengths in the middle of the visible part of the spectrum, the focal lengths of the spectrographs tended to be about 0.5 m, and grating groove densities between 600 and 2400 g/mm were typical. In addition to spectral resolution, the ability to measure Raman lines very close to the laser excitation is another measure of capability of an instrument. On some of these sys- tems Raman features shifted only 10 cm“ (or even less) from the laser line were possible to record. ‘Typical performance tests of one of these systems were published in 1985 [28] and 1987 [29]. Comparisons were made with a double I-m spectrograph in use at the time. Spectra recorded on a triple with a PMT had slightly degraded spectral resolution (relative toa scanning double), but the multichannel advantage implicit in the technique was clearly Copyright © 2001 by Taylor & Francis Group, LLC observed. At least on the microprobes (where the 2.5-mm-high pixels available in the intensified diode arrays at the time captured the entire available signal), the sensitivity of the individual detectors on the array was comparable to that of a PMT. So the time advantage to collect a spectrum was equivalent to the size of the array (ie., the number of pixels) [28,29]. Several of the triple spectrographs had additional capabilities built into them to provide fiexibility in function [27]. By reversing the sense of the dispersion of the first mono- chromator, the system could be converted to a “triple additive” spectrograph. When used in this mode, stray-light rejection was sacrificed (.e., low-frequency bands were no longer measurable with a multichannel detector), but subtle, stress-induced shifts of the bands along the Raman shift axis or changes in line shape could be measured. The usefulness for studying these subtle spectral effects was reported on scanning doubles in the early 1980s [30], but measurements on an additive triple provided higher precision in the mea- surement (by referencing to atomic lines, with the drive system fixed), and less likelihood for laser-induced changes during the measurement [31]. The triple spectrographs also had a “bypass” mode where samples could be measured through the third stage alone, which provided more signal. This last capability became especially useful after the holographic filters became available (see below). Based on this discussion, one might think that the double additive monochromators would haye no interest for use with multichannel detectors. This would follow from the fact that the high dispersion results in low “spectral coverage” on a multichannel detector. In fact, this spectrograph has become the instrument of choice for studying very subtle spectral changes when a system is dedicated to a particular application. In this case, a double additive used for this application has similar dispersion to a triple additive (with somewhat shorter focal length)—with the important difference that there is more light transmitted by the double. A major application has been the study of local strains in integrated circuits [30,31]. In silicon, the local strains affect electronic mobility and device performance; measuring the frequency of the Raman band can quantitate these strains. The integrated-circuit industry has employed Raman microprobes based on double mono- chromators and multichannel detectors for monitoring and mapping strain in silicon [30,31]. However, this gets ahead of the story somewhat, because the microprobe has not yet been addressed. B. Multichannel Detectors The implementation of multichannel detectors in a variety of spectroscopic applications was covered in a 1978 ACS symposium already cited in Ref. 26. In the opening chapter [32], Talmi reviewed the technology of these detectors at that time period. He noted that in attempting to reduce the time collection of spectroscopic data, both interferometric (Fourier transform or FT) and dispersive multichannel techniques were being implemented. ‘The FT method was having great success in the infrared part of the spectrum where the signal-to-noise ratio was limited by the detector noise. However, in the visible part of the spectrum, detectors are shot-noise limited, which precludes the analogous improvement when implementing this technology there. In fact, the multiplex disadvantage had already been observed. (The multiplex disadvantage refers to the noise that is spread over the whole spectrum because each point in the interferogram has signal from every wave- Iength.) What was important at the time was the recognition that the parallel detection of a multichannel detector provides observation of many wavelengths simultaneously without the potential for the multiplex disadvantage. Copyright © 2001 by Taylor & Franeis Group, LLC Whereas dispersive instruments with multichannel detectors have evolved into easy- to-use tools, they are limited to wavelength ranges in which good multichannel detectors are available. Because the long-wavelength limit of any of these detectors is the bandgap of silicon (~1.1 pum) or the response of the photocathode of an image intensifier, this, technology would not be of any use if a YAG laser operating at 1.06 nm was used as a source of Raman spectra. This was one of the reasons stimulating the development of FT techniques after it was realized that a long-wavelength laser eliminated the contaminating fluorescent background from many organic samples [33]. However, it is still possible to record at least part of the Raman spectrum with a charge-coupled devise (CCD), if the spectrum is recorded on the anti-Stokes side [34,35]. The obvious implication to this discussion that when multichannel detectors based on material sensitive to NIR radiation become available, it will be possible to build a dispersive, multichannel instrument for excitation wavelengths longer than 1 zm. Development of such detectors has been di cussed [36]. Again, we are getting ahead of ourselves because of the convolution of the many factors that determine the optimization of Raman equipment. By 1978 when the ACS symposium on multichannel detectors was convened [26], it was recognized that a multichannel spectrograph could provide large reductions in the time for collection of spectra. No one would use photographic plates anymore because of their low quantum efficiency (QE), the low accuracy and precision of the measurement, the tedious calibration, and especially the slow plate-development time. The time had come for introducing state-of-the-art TV cameras as well as other opto-electronic devices as spectrometric parallel detectors Any of these detectors included three essential componen + A transducer to convert a photon image to an electrical analog + A storage device to hold the latent electrical image + Areal-time readout of the stored image to a display monitor or to an A/D converter for digital storage ‘What is clear is that this type of device is ideally connected to a computer for storage and data processing. As computer technology evolved over the same time period, computer power increased and costs decreased, further fueling instrument evolution and more wide- spread acquisition of equipment. 1. Intensified Photodiode Array According to Talmi [26], although charge coupled devices (CCDs) and charge-injection devices (CIDs) were available in 1978, their performance and manufacturing yields limited their implementation as commercial products. Consequently, it was the monolithic silicon vidicon composed of photodiodes that was implemented most often. All the diodes on the device have a common cathode, and the isolated anodes are selectively addressed with a scanning, readout beam. After irradiation with photons between the ultraviolet (UV) and near IR, electron-hole pairs are produced and stored until readout. The scanning electron beam recharges the depleted surface of the photodiodes to a preset reversed-bias potential, and the amount of charge required to recharge is related to the photon intensity that previously illuminated the pixel. Because the readout noise of the device is about 2000 electrons rms (root-mean-squared), the smallest signal that could be detected was several thousand photoelectrons. The QE of the device was of the order of 10-80%. In order to achieve single-photoelectron sensitivity, an image intensifier was typically added for Ra- man applications. The amplification of the image intensifier provided a gain of about 1500, Copyright © 2001 by Taylor & Francis Group, LLC which boosted the capacity of the silicon photodiode to detect very weak signals. The addition of the image intensifier to the front end of the detector changed the wavelength response of the device to that of the photocathode, which was made of materials similar to those in a photomultiplier. Although photodiode arrays were available in two-dimen- sional format, commercial devices for Raman systems were typically one dimensional. The intensified diode array provided an additional advantage, which was the ability to “gate” the image intensifier, thus providing time-resolved spectroscopy. This was achieved by applying a bias voltage to the image intensifier, during which time the pho- toelectrons would not be accelerated to the back surface of the image intensifier mounted in proximity to the silicon array. When it was desirable to sense the Raman signal, the bias was turned off and the signal was collected. Depending on the quality of the image fier, the gating could be as fast as a few nanoseconds. On the downside, IPDAs are limited by the noise inherent in their design and by the deleterious effects of aging on the intensifier. Consequently, when the performance of CCDs improved and the chips became reliably available, they replaced the IPDA as the detector of choice for most Raman systems, 2. Charge-Coupled Device Charge-coupled devices appeared as reliable, commercial products for spectroscopic ap- plications in the early 1990s, The basic operation of this device is somewhat different from a photodiode array. The photodiode array is based on bipolar technology. ‘The pho- tosensing clement is a p-n junction that functions as a photodiode. Readout is directly through a FET (field-effect transistor). The CCD is fabricated on MOS (metal—oxide— semiconductor) technology. The charge is stored in the capacitor gate, Readout occurs after sequential charge transfer in what is termed the “bucket brigade,” where clocking of voltages on the electrodes controls the transfer of the photoelectrons. Because of the characteristics of the electronics, especially the low spontaneous generation of electrons, the CCD is inherently capable of detecting much smaller signals than PDAs, even though the dynamic ranges tend to be similar. There are numerous texts treating CCD design, one of which is included for the reader’s reference [37]. ‘As with the photodiode arrays, the CCD architecture has three functions: + Photon detection and charge collection + Charge transfer + Conversion of charge into a measurable voltage that is subsequently digitized and transferred to the control computer The wavelength response of a CCD depends, to first approximation, on the bandgap of silicon, just as with a photodiode array. Note that silicon is an “indirect-gap” material which means that the wavelength turn-on at 1.1 eV (1.13 nm) is not sharp. Because the absorption coefficient decreases at long wavelengths, the effective QE will also drop. Effectively, the long-wavelength response depends on the thickness of the well versus the electron-diffusion length. In addition, the transmission of any materials that have to trans- mit the signals before light hits the silicon well will affect the wavelength-dependent response, Modifying the device structure has been used to improve the wavelength char- acteristics, which consequently improves the QE. Because the gate is made of polysilicon that tends to absorb light at wavelengths shorter than 600 nm, thinning the gate improves the response in the middle of the visible range. The UV response can be enhanced by applying a UV fluorescent phosphor to the surface. Excitation of these phosphors with Copyright © 2001 by Taylor & Franeis Group, LLC wavelengths between 120 and 450 nm produces emission between 540 and 580 nm, which is in the middle of the peak response of the standard CCD. Other techniques that have been used include thinning the entire device, enabling illumination of the back side, in- creasing the depletion depth at the surface (i.e., the depth of the well), and using open polysilicon electrodes. + Back-thinning: Iumination of the array from the back of the device eliminates the problem of absorption by the polysilicon gates. However, because the film thi ness is comparable to the far-red wavelengths, there is often etaloning. Etaloning refers to coherent multiple reflections in the silicon, which results in ripples in the wavelength response, a phenomenon that is considered unacceptable for a Raman system. + Open polysilicon: About 33% of the polysilicon electrode is removed in the center of the pixel, as another means to eliminate the absorption by the polysilicon. + Deep depletion: Increasing the depleted volume near the surface increases the depth of the well. Because the long-wavelength photons travel farther before being ab- sorbed, the QE in the far red is enhanced. Because CCDs are used for many, large-volume applications, new devices are contin- ually appearing on the market, Pixel sizes as small as 10 um are available and provide higher spectral resolution, or superior band-shape analysis when that is an issue. [For routine applications, however, it is useful to remember that decreasing the pixel size while increasing the number of pixels on a chip of a given size will impose a load on computer memory and will increase the amount of time to read out the chip.] IV. HOLOGRAPHIC NOTCH FILTERS AND SINGLE SPECTROGRAPHS The introduction of the holographic notch filters to suppress stray light and the Rayleigh line [38] has been the most important innovation in simplifying dispersive Raman instru- mentation. The reduction of size and complexity potentiated by this development enabled use of Raman equipment by analysts with neither the time nor the inclination to be in- strumentation specialists. The first filters to appear were edge filters which passed Raman light to the long-wavelength side of the excitation (i.e., the Stokes Raman shift). Within a short period of time, notches with a narrower FWHM (15 nm, or 600 cm” in the green) appeared, allowing the recording of Stokes and anti-Stokes spectra simultaneously (39). Filters with FWHMs of half that size are currently available for most wavelength: The technology of these filters will certainly be covered more completely in a later chapter in this book, so details will not be covered here. However, it is impossible to overemphasize the importance of this innovation. The double subtractive premonochro- mator that functioned as a laser filter in triple systems became necessary only in cases when it was required to acquire signals closer to the laser line than the holographic filter will allow (effectively closer than 75~100 cm“), or when the laser excitation was tuned. Consequently, elimination of the filter stage immediately decreased the instrument size. It also became clear that the focal length of spectrographs did not need to be 0.5 m or larger, especially when the spectra were to be excited with a red laser. It is probably useful to describe the effect of the laser wavelength on the dispersion. Although the dispersion of a grating-based spectrograph is more or less constant in wave- Copyright © 2001 by Taylor & Francis Group, LLC length, the number of cm~'/am changes dramatically in different parts of the spectrum, as indicated in the following table: a 300 nm 400 nm 500mm 600 nm 700nm = ——-800 nm Disp tiem VA 6.25em VA 4em VA 28cm YA Dem VA 1.6 em YA Thus, a spectrograph with only a 0.25-m focal length can actually deliver about 1 em” in a 25-ym pixel in the red part of the spectrum! Another practical consequence of the elimination of the double subtractive filter stage is the relaxation in the requirements for design, alignment, and maintenance of Raman systems, It is probably not well understood how critical the wavelength registry is to the performance of a system composed of serial spectrographs. Every stage has to pass exactly the same central wavelength, and that registry has to be maintained as the wavelength is changed. If a Raman system is built with a single spectrograph, any issues of apparent “throughput” that can be so sensitively affected by poor registry between stages disap- pears. In a single-stage system, if the spectrograph is inadequately calibrated, the calibra- tion can be corrected ex post facto. With the availability of holographic notch filters and the reduced cost and complexity of Raman equipment, one sees the appearance of new applications to many analytical measurements and the yearly introduction of new instruments specialized for making par- ticular measurements V. FOURIER TRANSFORM-RAMAN (FT-RAMAN) INSTRUMENTS A. Background ‘The first FT-Raman spectra reported were made in 1984. Weber et al. measured the Raman spectra of gases using an argon laser and the interferometer at the Kitt observatory [40,41]. However, the analytical chemistry community missed the importance of these measure- ments. In the introductory chapter to their text, Chase and Rabolt [42] note that Chantry et al. [43], as early as 1964, suggested the use of infrared radiation for excitation of Raman signal with an interferometer for a detector. They demonstrated the feasibility of the tech- nique, with the important remark that the long excitation wavelength would avoid the excitation of fluorescence! However, the equipment at the time was not adequate to pro- duce good quality spectra, so the technique lay dormant until 1984. Following a discussion between Bruce Chase and Tomas Hirschfeld, existing Fourier transform infrared (FTIR) instruments were modified to measure Raman spectra generated with krypton and YAG lasers (emitting at 647.1 nm and 1.064 ym). The Raman signals were directed into interferometers (modified for detection in the far-red and near-IR part of the spectrum) at the emission ports [45-47]. This technology generated immediate interest in the community of analytical chemists and spectroscopists, especially those handling organic materials. The long laser wavelength of the YAG laser avoided the excitation of fluorescence that had plagued the measurements of many organic materials over the years. [The fluorescence interference can be significant even if the fluorescer is a low-level impurity because fluorescent efficiencies are typically many orders of magnitude larger than Raman efficiencies.] A second advantage provided Copyright © 2001 by Taylor & Franeis Group, LLC by FT-Raman equipment, also resulting from NIR lasers, is the excitation of spectra with no resonant enhancement factors [47]. This means that the band intensities in the spectra are more likely to be representative of the concentrations of the chemical species. Another factor affecting the rapid acceptance of the FT-Raman technology was the availability of many fully mature software functions for spectral manipulations that had already been developed for FTIR instrumentation. The software contained well-developed capabilities that were easily extended to Raman spectra. It is probably not an exaggeration to say that acquisition and use of FT-Raman equipment has been exponential. A full review of FT-Raman technology will not be attempted here; only features and impact on the Raman community will be covered. salient B. Characterisitics of FT-Raman Instruments The characteristics of interferometers for spectroscopic applications have been covered in excellent texts, one of the earliest of which is listed in the references [48]. Interferometers can effectively replace dispersive instruments when the noise of the measurement is gen- crated in the detector and is independent of the signal power. This is the case in the NIR and mid-IR (MidIR) part of the electromagnetic spectrum, where many detectors produce noise independent of the source and this noise dominates the interferometric signal. Under these conditions, the instrument is working in a regime of the Fellget or multiplex advan- tage. This means that signal from the entire spectrum is being collected simultaneously. [In contrast, in the visible part of the spectrum, both PMTs and CCDs measure signals that are essentially “shot-noise” limited, meaning that one sees the result of measuring statistical events (i.e., photon signals)—with the result that quadrupling the integration time doubles the signal-to-noise ratio (SNR).] In recent years, detectors with better noise characteristics (i.e., approaching the shot- noise limit) have been developed for FT-Raman systems. In these systems multiplexing can become a disadvantage because each spectral element contains shot noise from the sum of the entire spectrum, However, even under these conditions, an interferometer still provides the luminosity (tendue), or Jacquinot advantage, over a dispersive system. [Jac- quinot had shown that because of the absence of an entrance slit in an interferometer, for equivalent spectral resolution, the luminosity of an interferometer is greater than a di persive spectrometer.] One remaining characteristic of FI-Raman systems that should be mentioned is the wavelength, or Cones, advantage. In order to keep track of the path difference of the two arms of an interferometer, a HeNe laser beam is used collinearly with the light path, and fringes are counted. Because the wavelength of the HeNe laser is known to a very high precision and accuracy, the wavelength accuracy of the entire spectrum produced by the interferometer is quite high. This makes the comparison of spectra very accurate and incteases the information provided when subtracting a spectrum of a known species from the spectrum of a sum of components. A property of organic samples that can also affect the results of FT-Raman measure- ments is the reabsorption of the Raman light by the sample itself. This arises from the overlap between the Raman emission and the direct absorption by vibrational overtones and combinations that occur between 1,064 jum (the excitation) and 1.786 pm (the end of the Raman spectrum at the 3800-cm™ shift from the laser). These absorptions will affect relative intensities in the Raman spectrum, especially in large samples where the laser is penetrating deep below the surface. The effects of self-absorption will be especially 2001 by Taylor & Francis Group, LLC important when extracting quantitative information from the spectra. In addition, when samples are highly scattering, the optical path length will not be reproducible, even if the sample placement is precise. Such differences in optical path lengths will make corrections for self absorption close to impossible. These phenomena have been discussed and dem- onstrated [49]. In addition, these same problems can exist in spectra acquired in the visible region on dispersive systems when the samples are absorbing due to electronic transitions. ‘The following is a summary of the relative advantages and disadvantages of FT-Raman equipment versus dispersive equipment: Advantages + Long-wavelength avoidance of fluorescence (especi ly for organic samples) + Nonresonant conditions + Interferometer properties + Fellget’s (multiplex) advantage: Collect signals from all wavelengths simultaneously * Btendue (Jacquinot) advantage: Amount of signal detected by interferometer is determined by entrance aperture which does not have to be closed to im- prove spectral resolution as in a dispersive system + Connes advantage: Wavelength accuracy and precision is extremely good be- cause of the use of a HeNe laser to reference the interferometer Disadvantages * Overall loss of signal due to the A‘ dependence of the scattering process + Multiplex disadvantage, which occurs because signal is being detected at all wavelengths simultaneously so shot noise from the entire spectrum is spread into every spectral element * Self-absorption of the Raman light generated by a NIR laser VI. RAMAN MICROSCOPY A. Introduction Anyone who has leamed to use Raman spectroscopy before the introduction of the mi- croprobe, and then has been afforded the opportunity to take advantage of this sampling device, appreciates the enormous power provided. To dismiss the usefulness of a micro- probe by claiming that sample quantity is not limited indicates a lack of understanding of the performance of the device. In order to achieve high spatial resolution, approaching the diffraction limit, one needs high-numerical-aperture (n.a.) focusing optics [50]. A laser beam focused by an objective will have a beam diameter described by something between a focused Gaussian (when the lens diameter >> beam diameter) and an Airy disk (where a beam of uniform intensity fills a lens). The expressions for these quantities are respectively Gaussian: 4a ma. Wy (ie., effective na.) Copyright © 2001 by Taylor & Franeis Group, LLC na where « represents the beam waist defined as the diameter at the first node. Note that in spite of the fact that these two expressions describe very different conditions, the values produced by the two equations are quite similar. What one sees is that in order to achieve high spatial resolution (which means a small beam spot), a high-numerical-aperture optic is required. If the 180° scattering scheme is used, then the additional advantage of col- lecting light in a very large cone becomes a free payoff! Finally, when this small spot from which Raman light is collected is imaged onto the entrance slit, the image size can be smaller than 100 jm. That means that all the collected light will pass to the detector if the monochromator entrance slit is 100 jum or larger. [The Jacquinot advantage of an FT-Raman system effectively disappears when comparing an FT system to a dispersive microprobe!] The end result of implementing high-n.a. optics for achieving microspots is larger collection efficiencies and higher throughput into the spectrograph. The discussions that follow will indicate how these systems also discriminated against light originating, from material other than the microsample (‘“confocality”) providing additional benefit. ‘The experience that we have had in this laboratory over the last 21 years has been that unless there is an important reason to sample large volumes (which clearly exists in certain cases), the use of a Raman microscope/microprobe provides easier-to-acquire, higher-qual- ity spectra According to Delhaye and Dhamelincourt [51], the earliest descriptions for using a microscope as a Raman sampling device appeared in 1969 [52,53]. In 1973, Tomas Hirsch- feld (the same Tomas Hirschfeld who was influential in the development of FT-Raman) published an abstract describing the basic requirements for a Raman microprobe [54]. The concept was first demonstrated at the IVth International Conference on Raman Spectros- copy in 1974 in Brunswick, Maine [55,56]. At that meeting, the group at the National Bureau of Standards (NBS) (now NIST) demonstrated the ability to acquire spectra from microsamples, whereas the group at Lille demonstrated the ability to acquire Raman images B. Raman Imaging In their 1975 publication [51], Delhaye and Dhamelincourt clearly describe the three meth- ‘ods that could be used to acquire Raman images of samples (see Fig. 1). They show two schemes described as microprobes, and a third described as a microscope. The various methods to acquire images involve different means to illuminate samples and different types of detector, Because of how important Raman microscopy has become, it is useful to describe these methods here: + Microprobe with Single-Channel Detector: In this scheme, the laser is focused to a diffraction-limited spot, and then rapidly deflected to illuminate a line on the sample. The light scattered by the illuminated line of sample is imaged onto the entrance slit of a monochromator whose grating is positioned to present a Raman line of interest to a single-channel detector. The output of the detector is transferred to an oscilloscope and synchronized with the scanning on the sample. The second dimension of the image is produced by translating the sample in a direction at right angles to the line focus. Copyright © 2001 by Taylor & Francis Group, LLC Point Focus Line Focus | Global Monochromator Monochromator llumination Single Channel MultiChannel Grating Filter Detector Detector Multichannel Detector Focused Line Laser Beam Huination L Global IMlumination Line Frame Frame Only NO SCANNING SCAN Figure 1 Raman imaging schemes, as laid out by Delhaye and Dhamelincourt [55] + Microprobe with Multichannel Detection: In this scheme, the sample is again il- luminated with a line focus, the illuminated area is imaged onto the entrance slit of the monochromator, and the monochromator is set to a Raman line of interest. One axis of the multichannel detector is oriented parallel to the slit direction, providing simultaneous detection of spectra from each point on the line. Each row of the detector monitors a different point on the illuminated line, providing that the monochromator is “stigmatic” (i.e., there is little optical blurring by the mono- chromator at the focal plane). As in the first scheme, the second dimension is achieved by translating the sample in a direction at right angles to the line focus + Microscope with Multichannel Detection: In this scheme, an unfocused laser beam. illuminates the area of the sample. An optical filter isolates the selected Raman Tine, and the multichannel detector produces a “picture” of the sample. At the time ince of this publication, the optical filter was a monochromator used as a filter. [ then, this same concept has been implemented with dielectric filters [57], AOTFs (acousto-optic tuning filter) [58], and LCTFs (liquid-crystal tuning filter) [59,60]; these technologies will certainly be covered in later chapters on microscopy and imaging. Each of the methods laid out by Delhaye and Dhamelincourt has its advantages and disadvantages. The first two methods are based on microprobing (point illumination) and, as such, can be confocal, a concept that was implemented some time in the 1980s and will be described below. The second method multiplexes sample illumination in one di mension, which means that there is some time advantage over the first, when there is sufficient laser power available so that the illuminating intensity is not a limitation. The Copyright © 2001 by Taylor & Franeis Group, LLC third method multiplexes the signal acquisition in two dimensions and, as such, has the potential to acquire an image much more rapidly than either of the former methods. How- ever, the lack of ability to prevent “out-of-focus” radiation from reaching the detector means that the images can be overwhelmed by Raman and fluorescence signals generated away from the focal plane. When the issue of confocality is treated, this will become clearer. C. Raman Microprobe Following the abstracts and presentations that appeared in the 1974 International Confer- ence on Raman Spectroscopy, Dhamelincourt [61] and Rosasco {62] systematically laid out the performance criteria effecting the microprobe design in 1979 and 1980. The system that was implemented at the NBS [62] used a specially designed sampling optic—a con- caye mirror formed as an ellipse in which the sample was placed at one focus and an exit pinhole at the second focus (see Fig. 2). This “spatial filter” at the exit focus was reimaged on the entrance slit of the monochromator and was meant to minimize the signal contri- butions from everything except the sample at the focus. In his Ph.D. thesis [61], Dhamelincourt detailed optical considerations for the design and implementation of a Raman microprobe and microscope, based on commercial optical microscopes, a design that was commercialized as the MOLE™ (Molecular Optical Laser Examiner) by Lirinord/Jobin Yvon. In this system, a beam splitter in a standard epi- illuminator (Fig. 3) was used to disentangle the incident laser beam from the Raman scattered beam, The microscope optics are infinity focused. When functioning as a microprobe, optics between the microscope and the monochromator or spectrograph were placed to transfer the image of the point focus at the sample onto the entrance slit and to transfer an image of the back aperture of the microscope objective onto the grating. Acquisition of first- generation microprobe spectra was done with a PMT, which was slow. When functioning Laser Entry | 7 | To Monochromator Elliptical Sample Reflector Plane Image Plane Figure 2 Sampling scheme of the NBS (NIST) microprobe. (From Ref. 62.) Copyright © 2001 by Taylor & Francis Group, LLC To Monochromator Raman Return Laser Entry Beam Splitter ‘Sample Plane Figure 3. Epi-illumination scheme for using a metallographic microscope with beam splitter to build a Raman microprobe. (From Ref. 61.) as a Raman microscope, the sample was illuminated globally and was imaged onto the grating through a wide entrance slit, The system maintained the optical resolution provided by an optical microscope in both microprobe and microscope modes. In microscope mode, the angle of the grating determined which Raman band was selected to be imaged onto the multichannel detector that was functioning as a camera. The monochromator used in this system was a double I-m system based on concave holographic gratings. Images were acquired with SIT (silicon-intensified target) or SEC (secondary emitting cathode) multi- channel detectors. The instrument described in this work was commercialized as the MOLE. Eventually, as multichannel detectors improved and the spectrographs were redesigned to take advantage of the new possibilities, these new designs were incorporated into newer Raman microprobes. Some of these new systems also incorporated the concepts of con- focality in both microprobing and imaging. Because of the importance of confocality, it will be treated in more detail below. The background of the concurrent development of the two microprobes that were realized in the United States and France is interesting. The NBS system was developed under funding by the U.S. Air Force whose motivation was to follow the chemistry of radioactive microparticles released into the atmosphere by the atomic tests of the 1950s. Hirschfeld [63] pointed out that only a Raman microprobe could provide the information on chemical speciation necessary to understand the environmental fate of these materials. In France, the group in Lille, under the direction of Professor Michel Dethaye, was creating a molecular microprobe that would provide chemical information complementary to that provided by Castaing’s (elemental) electron microprobe. In both locations, it was recognized that the instrumentation was an enabling tech- nology for microanalysis of all kinds. Examples of applications of the technology included the study of polymers, mineralogy and gemology (both solid phases and fluid inclusions were analyzed), and contaminant identification in a variety of manufacturing environments, catalysts, atmospheric pollutants, histology [61,64]. Copyright © 2001 by Taylor & Franeis Group, LLC To anyone with training in physics or materials science rather than chemistry, it was also clear that the instrumentation had enormous potential for studying crystalline mate- rials; that is, the crystalline phase had information as valuable to materials scientists as the chemical composition had to chemists. We have the following examples: + The solid-state chemistry changes occurring in ceramics during processing can easily be followed with a Raman microprobe [65]. + The potential to follow strains in silicon on a size scale commensurate with device structures was published as early as 1983 [30,31]. Quantification of local stress is important for the integrity of integrated circuits (¢.g., delamination of materials due to stress can develop during processing) and for control of electronic mobility (which determines the electrical properties of the micro-devices). In more recent years, local stress has been mapped more systematically in devices as an aid to engineering design and to control of the manufacturing processes {66] + In individual 7-~m polymer fibers, independent orientation and crystallization phe- nomena occurring during the spinning and drawing process were measured [67]. D. Confocal Raman Microscopy and Confocal Imaging In 1993, Barbillat, et al. [68] summarized the various methods for imaging, including some instrumental developments that enabled confocal mapping. The principle of confo- cality, as will be described below, limits the amount of radiation that is collected from out-of-focus regions of the sample. Barbillat’s analysis was done on Dilor’s XY multistage spectrograph system. However, there is nothing specific to the design that depends on the type of spectrograph employed. Consequently, as single-stage instruments were adapted for light throughput reasons, the same principles were implemented on the newer systems. During the period of the mid-1990s, direct-image formation of Raman light through various filtering arrangements reappeared. Because holographic notch filters effectively eliminated the unshifted laser signal, simple, compact filtering devices could be used, and some novel solutions appeared. The Raman line could be selected with an interference filter [57,69], an AOTF [58] and with an LCTF [59,60]. These technologies will certainly be covered in more detail in later chapters. Because a conventional microscope has limited capabilites for axial spatial resolution, these global Raman images do not represent thin sample planes. Images specific to thin sample planes can be regenerated by mathematical manipulation of a series of digital images collected as a function of depth. The group at the University of Michigan under the direction of Professor Michael Morris implemented constrained iterative image-resto- ration techniques to generate a stack of well-resolved Raman images describing the three- dimensional topology of the samples [70]. To generate these images, an instrument would have global illumination with direct Raman imaging as a means to collect a series of images at a selected wavelength and as a function of depth. This method has been shown to successfully produce three-dimensional images. However, there is some remaining noise in any one image due to the multiplexed method of image collection. 1. Confocal Principle By extending the principles of confocal microscopy [71], it became clear that it is possible to collect interference-free Raman maps of small samples in matrices, even if these ma- tices are strong Raman scatterers or fluorescers. 2001 by Taylor & Francis Group, LLC Both carly microprobe designs (NBS and Lille) have the capability for closing exit apertures in planes conjugated to the sample. The description of both instruments shows recognition of the value of these pinholes in restricting the sample zone from which Raman radiation is collected. In the NBS system, this aperture (shown in Fig. 2) is called a spatial filter, and in the Lille design, it is an exit pinhole placed at an image plane behind the microscope. In both cases, the aperture is subsequently imaged onto the monochromator entrance slit. These techniques are actually equivalent to those of scanning confocal mi- croscopy that has been developed to increase image quality and contrast in optical mi- croscopy [71]. Both groups studied the ability of the exit apertures to reject out-of-focus radiation [61,62]. The principle of confocal Raman microscopy is illustrated in Fig. 4 (reproduced from Ref. 68). In the center of the figure, Raman light is collected from the plane in which the laser is focused. That Raman light will be refocused through the pinhole in the plane conjugated to the sample and will be efficiently detected by the photodetector (which includes the Raman spectrograph). On the right side of the figure, Raman light is generated from material above the focal plane (by an amount equal to +Az), even though the laser is not in focus in that plane. However, its detectivity is decreased because the signal will not be effectively transmitted by the conjugate pinhole. The left side of the figure shows the analogous case for material below the focal plane. (Note that if the sample is opaque, the laser will not penetrate below the focal plane, and the out-of-focus signal will only be generated in the medium above the sample.) The top of the figure shows the behavior of the detected signal as a function of distance from the focal plane. The FWHM of the intensity profile will depend on the n.a. of the objective and the size of the pinhole. Studies of the lateral and depth resolution show significant improvement over a conventional microscope [68,71]. Discrimination Lateral resolution Depth Conventional Microscope (22) (sae) v wl 2)" ‘ Confocal Microscope ( 2) (ae) where v= (27/A)r sin ais a normalized distance perpendicular to the optic axis, J; is the first Bessel function, and u = (27/A)z sin’a is a normalized distance from the focal plane of the microscope objective. Figure 5 shows the results of a depth profile of a thin slab. The sample was a 0.5 um film of polystyrene and the plot on the left side of the figure shows the intensity of the strong band at 998 cm” as a function of sample position relative to the microscope focus. The FWHM of the depth profile is about 2 jum. The ability to reject out-of-focus radiation from the collected Raman signal provides the following advantages: * Discrimination of well-defined spatial region in a complex multiphase specimen + Rejection of stray light Copyright © 2001 by Taylor & Franeis Group, LLC Iz PHOTODETECTOR SIGNAL, -Az + hz 0 Z_ AXIAL DISTANCE PHOTODETECTOR CONJUGATED PINHOLE DIAPHRAGMS BEAM SPLITTER LASER BEAM NN MICROSCOPE, OBJECTIVE FOCUS PLANE —» ~ Z:0 ~ SAMPLE THIN SLICE Figure 4 Principle of confocal microscopy and depth discrimination. (From Ref. 68.) © 2001 by Taylor & Frcs Grow, LLC Raman Intensity @998 cm? (arbitrary units) 10 drwy Polystyrene + dz=0.5 pm -10 ‘ZL (um) Figure 5 Depth profile of a 0.5-jum slab of polystyrene. (From Ref. 68.) * Rejection of matrix signal from the signal of a small embedded inclusion + Rejection of fluorescence The last point bears some clarification, It has been observed that spectra of many materials that normally fluoresce so much that it is impossible to record their Raman spectra can be recorded when examined in a microprobe [72]. This is a result of several factors: + Nonfluorescent regions can be identified, if the material to be examined is larger than a few cubic micrometers. + The high laser flux in the sampling volume can effectively quench the fluorescence in a few minutes (rather than a few hours, as was necessary in macro-devices). + Closing the confocal aperture further reduces the fluorescence contribution o1 nating outside of the primary focal volume. This works because, it is believed that some of the fluorescence is emitted away from the point of illumination following exciton migration of the fluorescent excited state. 2. Confocal Raman Mapping Clearly, if one rasters the sample under the focused beam, it is possible to acquire a two- dimensional array of spectra from which a confocal Raman map can be reconstructed. With a two-dimensional detector, this process can be doubly multiplexed; that is, in ad- dition to spectral multiplexing along the long dimension of the detector, it is, in principle, possible to multiplex in one spatial direction along the short dimension of the detector. This is achieved by illuminating a line that is projected onto the entrance slit of the spectrograph. The original implementation of line illumination was not confocal. It involved a line focus on the sample [73,74]. If the line is illuminated by rastering the laser beam, there is a unique method to project this line focus onto the sample without introducing aber- rations in the focus [75]. In this scheme, illustrated in Fig. 6, a scanning mirror in the laser entry path, in combination with a focusing lens, produces a line focus in the focal plane of the microscope. Because the scanning mirror is optically conjugated with the exit pupil of the microscope objective, the mirror is imaged on the exit aperture of the objective and the laser beam then always illuminates the full pupil. The microscope objective pro- Copyright © 2001 by Taylor & Franeis Group, LLC CONCAVE GRATING coy prrectoR STIGMATIC ‘SPECTROGRAPH Figure 6 Confocal line scanning illumination. (From Ref. 68.) duces a diffraction-limited line focus on the sample in the microscope object plane. Figure 7 shows the Raman light returning through the same path; the scanning mirror returns the Raman light from each sample point to the undeviated path. Consequently, Raman signals from every point on the sample can be filtered by the confocal pinhole before entering the spectrograph. The author/inventors of this technique named it CORALIS for confocal Raman line scanning [68]. For a comparison of the performance of a global imaging system to a system pro- ducing Raman maps created from a confocal line scan, the reader is referred to Ref. 76. © 2001 by Taylor & Francis Group, LLC 0° 5 Ss ms Figure 7 Raman maps of a multiphase fiuid inclusion within a quartz host produced by confocal line scanning. Left side, top: water phase: right side, top: CO,; bottom: nacholite or NaHCO, In this case, the samples studied were polymeric; they were transparent and exhibited significant amounts of fluorescence. The authors found that “as a consequence of the absence of depth resolution, the global illumination technique appeared more vulnerable to artifacts arising from scattering effects due to the sample geometry and fluorescence.” Vil. REMOTE SENSING For more than 15 years, there have been reports of remote measurements of Raman signals using fiber optics to deliver the laser beam to the sample and to collect the Raman signal and deliver it to the spectrograph. These optrodes provide convenience to the researcher to examine samples inconvenient to measure close to an instrument. Following chemical reactions is one important class of such samples. Once it was realized that one could follow chemical reactions through Raman spectra, interest in monitoring manufacturing processes followed. This topic will certainly be covered in a later chapter. Here, the ev- olution of the design of the probe, an enabling technology, will be reviewed. Clearly, the uses for these probes has been the driving force in their development. For a survey of Copyright © 2001 by Taylor & Franeis Group, LLC instrumentation for monitoring and control of chemical manufacturing, and quality control/ quality assurance (QC/QA) applications, the reader is referred to the relevant chapter. There are essentially two classes of design. In the earliest designs, the probe w: inserted into the sample: this class of probe has been termed “immersion,” “insertion,” “nonimaging,” and “n-around-1.” In the second type of probe, there is a focusing element and the sample can be examined “through a window.” This class of probe has been termed “imaging,” “coaxial,” and “focus,” depending on the literature reference. A. Nonimaging Insertion Probes In 1984, McCreery [7] reported the use of a fiber bundle, close-packed at the sampling end, and arranged in a slit configuration at the spectrograph entrance slit end. A central fiber delivered the laser beam to the sample (liquids), there was no focusing element, and the Raman signal was generated in the overlap of the projected cones of the fibers. A group of workers at Dow Chemical Company used a similar design, but with the fibers angled in order to increase the overlap volume [78]. Another variation on this design was produced at the Westinghouse, Savannah River facility [79]; the fibers were mounted parallel, but then polished at an angle, which had the same effect of increasing the overlap volume. These designs are shown schematically in Fig. 8. The various parameters determining the collection efficiencies in these designs were compared in several publications [80-82]. It was found that the size of the fibers, the distance between them, the fiber n.a.’s, and the angle between them (if angled) all affected the sampling volume and the depth of field. It was concluded that different probe designs would be required for examining clear liquids and solids, as compared to highly scattering sample media (turbid liquids and powdered solids). B. Imaging Focused Probes Carrabba and Rauh [83] produced a small, ruggedized, focused design for an insertion probe using two fibers. One fiber delivered the laser to the sample and the second fiber carried the Raman light; a beam splitter separated the two beams before the focusing lens Position |__ Position Position ‘Sample {- Sampie | ‘Sampie A excitation pathway + Collection pathway Figure 8 Schematic representation of the various 6 X 1 fiber-probe designs, Top: Cross-sectional view of laser delivery (center) and Raman collection fibers, at sampling end; bottom: Side view of fibers: (A) original MeCreery design; (B) Dow optimized design; (C) Savannah River chamfered design, Copyright © 2001 by Taylor & Francis Group, LLC at the sample end. In this design, there were already some optics for filtering the Raman signal from the fiber. The Kaiser [84] and Dilor [85] probes also use two fibers (one for laser delivery and one for Raman return) and there is a lens for focusing onto the sample. Both designs use a filter for the delivered laser beam and a holographic notch filter to remove the laser wavelength from the Raman return beam, The reader is referred to Chapter 3 by Tedesco et al. and to Ref. 86 by Lewis and Griffiths who have reviewed the development and design of the probes in more details. Vill. LASER EXCITATION WAVELENGTHS AND APPLICATION OPPORTUNITIES The multitude of laser wavelengths that are available today means that the choice of wavelength for a project is dictated by the sample properties rather than what instrumen- tation is available. Some of the considerations for wavelength choice have already been mentioned in the text, but it is probably useful to summarize what needs to be considered: + One of the general properties of Raman scattering is the A‘ dependence. This refers to fourth-order behavior of the scattering intensity which increases as the wave- length decreases (or equivalently, as the frequency increases) + Lasers generating light at particular wavelengths need to be coupled to systems with detectors sensitive to the Raman spectra being generated, Figure 9 shows the Raman range for some of the currently available laser wavelengths. For all visible and UV laser wavelengths, almost all systems are currently being used with mul- annel detectors, usually CCDs whose wavelength response can be tuned for the selected range. For spectra excited with YAG lasers, most systems are FT-Raman systems, with Ge or InGaAs single-channel detectors + Resonance phenomena in which the Raman intensity is increased can be accessed by selecting a laser wavelength which couples effectively to an electronic excitation (RR = resonance Raman) or to a noble metal resonance (SERS = surface-enhanced Raman scattering). Resonance can be a desired or a nondesired effect, depending on the goals of the measurement. In many cases, the YAG laser can be used to avoid resonance effects. * When considering the optical resonances of the sample to aid in the selection of laser wavelength, the self-absorption of the sample also needs to be considered. When exciting organic materials or aqueous solutions in the NIR, this can mean absorption due to overtones and combinations of vibrational fundamentals. The self-absorption will affect relative intensities, which can severely affect the ability for quantitation. + Laser absorption can cause other unwanted phenomena such as sample heating (especially in the solid form), photochemistry, and sample transformation. + A judicious choice of wavelengths will avoid the excitation of fluorescence. In the case of organic materials, this usually means either exciting in the NIR, as dis- cussed in Section V, or in the UV [87]. Asher [87] has pointed out that if the excitation wavelength is short enough, the wavelength of the largest shifted Raman band will still be shorter than the start of the fluorescence emission. In the case of inorganic materials, there are often fluorescent emissions from transition metals or Copyright © 2001 by Taylor & Franeis Group, LLC NIR absorptions 9 tht a +} | A (nr 1064-1786400 725-18 617-258, e-em Tan EET, 05-4800 tation wavelengths, the Stokes Raman range (3800 em) displayed as a her psig ©2001 rare carth metals. In order to avoid these interferences, one needs to consider the details of the system being studied. However, it has been observed that the fluo- rescing impurities that tend to be present in glasses and ceramics pose much less of a hindrance when exciting in the blue to red part of the spectrum than when exciting in the near IR. IX. SUMMARY AND FUTURES It is difficult to say with certainty how Raman instrumentation will evolve over the next 10 years. However, there are several instrumentation developments that are beginning to appear as commercial products. These include more systems engineered for dedicated process and/or QA/QC applications, NIR multichannel detectors for use on a dispersive Raman instrument with long-wavelength lasers, UV microprobes, and near-field Raman microprobes that can complement atomic force microscopes (AFMs) or scanning tunneling microscopes (STMS), or their variants. It is an exciting time to be a Raman researcher. ACKNOWLEDGMENT The author wishes to thank all colleagues that were kind enough to take the time to confirm or correct her knowledge of the developments covered here. REFERENCES 1. CV Raman, KS Krishnan. A new type of secondary radiation, Nature 121:501; (31 March 1928) (cabled to Nature on 16 February) 2. A Jayaraman, AK Ramdas. Chandrasekhara Venkata Raman. Phys Today 57-64, 1988. 3. E Lommel, Wiedemanns. Ann D Phys 3:251, 1878. 4. A Smekal. Zur Quantentheorie der Dispersion. Naturwissenschafter 11:873, 1923. 5. HA Kramers, W Heisenberg. Z Phys 31-681, 1925. 6. E, Schrodinger. Ann Phys 81:109, 1926. 7. PAM Dirac. Proc Roy Soc London A 114:710, 1927. 8. CV Raman. On the molecular scattering of light in water and the colour of the sea. Proc Roy Soc London A 101:64, 1922 9. CV Raman, Indian J Phys 2:387, 1928. 10. CV Raman, A change of wavelength in light scattering. Nature 121:619, 21 April 1928 (sent to Nature on 8 March). 11, B Schrader, ed. Infrared and Raman Spectroscopy, Methods and Applications, Weinheim: VCH, 1995, 12, G Landsberg, L Mandelstam. Eine neue Erscheinung bei der Lichtzerstreuung in Krystallen, Na- turwissenschafien 16:557 (13 July 1928) (submitted 6 May). 13, AS Davydov. Theory of Molecular Excitons. M Kasha, M Oppenheimer, Jr (transl). New York: McGraw-Hill, 1962. (A comment about the use of “combinatorial scattering spectra” in the original Russian is noted on page v of the translators’ preface.) 14. WW Coblentz. Investigations of Infrared Spectra. Washington, DC: Camegie Institution, 1905. (re published by The Coblentz Society, Norwalk, CT, 1962). 15, Brandtmuller and Keifer, Spex Speaker (1978), produced for the 50th anniversary of the discovery of the Raman effect. 16. D Long. Early History of the Raman effect. Int Rev Phys Chem 7:314~349, 1988, 17. A Yariv. Introduction to Optical Electronics. New York: Holt, Rinehart and Winston. 18. G Herzberg. Molecular Spectra and Molecular Structure, Il. Infrared and Raman Spectra of Pola- tomic Molecules. New York: Van Nostrand Rinchold, 1944. (Although this text does not present a description of instrumentation, it is easy to find examples of photographic images of spectral lines and what appears to be densitometer traces of these images.) 19. CH Townes. In: JR Singer, ed. Advances in Quantum Electronics, New York: Columt Press, 1961, pp 3-11 Copyright © 2001 by Taylor & Franeis Group, LLC 20. 21. 22. 24, 25, 26. 21. 28. 30, 31 32, 34, 35, 36. 37. 38, 39. 40. 41 42. 43, 45, 46, 47, 48, 49. 50, SPS Porto, DL Wood. J Opt Soc Am 52:251, 1962. H Kogelnick, SPS Porto. J Opt Soc Am 53:1446, 1963. JA Konigstein, RG Smith. J Opt Soc Am 54:1061, 1964, M Leclercq. Etude de premonochromateurs soustractifs et de spectrometres Raman a reseaux hol- ographiques concaves. PhD thesis. L’Universite des Sciences et Techniques de Lille, 1975; M Le- clereg, F Wallart. Description of a zero-dispersion force-monochromator for low frequency Raman spectrometry. J Raman Spectrose 1:587-593, 1973. M Lectereq, B Sombret, F Wallart. CR Acad Sci Paris 177B:377, 1973. Burle Technologies. Photomultiplier Handbook. Burle Technologies, Inc., 1980, Y Tali, Multichannel Image Detectors. ACS Symposium Series 102. Washington, DC: American Chemical Society, 1979. A Deffontaine, M Bridoux, M Delhaye, E DaSilva, W Hug. The third generation of multichannel Raman spectrometers. Rev Phys Appl 19:415~421, 1984, R Grayzel, M Leclercq, F Adar, M Hutt, M Diem. An automated micro/macro Raman spectrograph system with multichannel and single-channel detectors in new molecular/crystalline microprobe. In: Microbeam Analysis—1985. San Francisco: San Francisco Press, 1985, pp 19-24. F Adar, J Lerner, Y Talmi. The signal-to-noise advantages of multichannel detectors on Raman microprobes. In: Microbeam Analysis—1987. San Francisco: San Francisco Press, 1987, pp 141— 143, P Zorabedian, F Adar. Measurement of local stress in laser-recrystallized lateral epitaxial silicon films over silicon dioxide using Raman scattering. Appl Phys Lett 43:177-179, 1983, T De Wolf, HE Maes, SK Jones. Stress measurements in silicon devices through Raman spectros- copy: Bridging the gap between theory and experiment. J Appl Phys 79:7148-7155, 1996. Y Talmi. Optoelectronic image detectors in chemistry, an Overview. In: Y Talmi, ed. Multichannel Image Detectors. ACS Symposium Series 102. Washington, DC: American Chemical Society, 1979. DB Chase, JF Rabolt. Fourier Transform Raman Spectroscopy. San Diego, CA: Academic Press, 1994, J Barbillat, E Da Silva, B Roussel. Demonstration of low-frequency performance and microanalysis capability of multi-channel Raman spectroscopy with near-infrared excitation. I. J Raman Spectrose 22:383-391, 1991. J Barbillat, E Da Silva, JL Hallaert. Multi-channel micro-Raman spectroscopy with excitation, Il, J Raman Spectrosc 24:53-62, 1993, D Chase. Arrays for detection beyond one micron. In: JV Sweedler, KL Ratzlaff, MB Denton, eds. Charge-Transfer Devices in Spectroscopy. Weinheim: VCH, 1994, chap 4. GC Holst. CCD Arrays, Camera, and Displays. Bellingham, WA: SPIE Optical Engineering Press, 1996. MM Carrabba, KM Spencer, € Rich, D Rauh. The utilization of a holographic bragg diffraction filter for Rayleigh line rejection in Raman spectroscopy. Appl Spectrosc 44:1558~1561, 1990. B Yang, M Morris, H Owen. Holographic notch filter for low-wavenumber stokes and anti-Stokes Raman spectroscopy. Appl Spectrosc 45:1533-1536, 1991 A Weber, DE Jennings, JW Brault, High resolution Raman spectroscopy of gases with a Fourier transform spectrometer. Proceedings IXth International Conference on Raman Spectroscopy, Tokyo, 1984, pp 58-61 DE Jennings, A Weber, JW Brault. Raman spectroscopy of gases with a Fourier transform spec- trometer: The spectrum of D2. Appl Optics 25:284-290, 1986. DB Chase, JF Rabolt. Fourier transform Raman spectroscopy: From concept to experiment, In: DB Chase, JF Rabolt, eds. Fourier Transform Raman Spectroscopy. San Diego, CA: Academic Press, 1994, GW Chantry, HA Gebbie, C Helsum, Interferometric Raman spectroscopy using infra-red excitation, Nature 203:1052, 1964. T Hirschfeld, B Chase. Appl Spectrosc 40:133, 1986. DJ Moffatt, H Buijs, WF Murphy. Appl Spectrosc 40:1079, 1986. CG Zimba, VM Hallmark, JD Swalen, JF Rabolt. Appl Spectrosc 41:721, 1987. YM Hallmark, R Sooriyakumaran, RD Miller, JF Rabolt. J Chem Phys 85:7413, 1986. J Chamberlain. The Principles of Interferometric Spectroscopy. Chichester: Wiley-Inters 1979, CJ Petty. Self-absorption in near-infrared Fourier transform Raman spectrometry. Vibr Spectrosc 2: 263-269, 1991 M Bom, E Wolf. Principles of Optics—Electromagnetic Theory of Propagation, Interference and Diffraction of Light, Oxford: Pergamon Press, 1975, frared 2001 by Taylor & Francis Group, LLC 61 62. 63, 64, 65. 66. 67. 68, 69, 10. 1 72. 73. 74. 1S. 16. 71, 8, 79, 80. 81 82, 83. M Delhaye, P Dhamelincourt. Raman microprobe and microscope with laser excitation. J Raman Spectrosc 3:33-43, 1975. SK Freeman, DO Landon. Anal Chem 41:398, 1969. M Deporeg, R Demol. Rev GAMS 3:214, 1969. T Hirschfeld. J Opt Soc Am 63:476, 1973. M Delhaye, P Dhamelincourt, IVth International Conference on Raman Spectroscopy, Brunswick, ME, 1974, GJ Rosasco, ES Etz, WA Cassatt. IVth International Conference on Raman Spectroscopy, Bruns- wick, ME, 1974. Renishaw, PLC. European Patent 0 543 578 AI (filed Nov. 13, 1992; published May 26, 1993). MD Schaeberle, JF Turner II, PJ Treado. Multiplexed acousto-optic tunable filter (AOTF) spectral imaging microscopy. Proc SPIE 2713, 1994. PI Treado, IW Levin, EN Lewis. Appl Spectrosc 46:1211, 1992. HR Morris, CC Hoyt, PJ Treado. Imaging spectrometers for fluorescence and Raman microscopy: Acousto-optic and liquid crystal tunable filters. Appl Spectrosc 48:857-865, 1994, P Dhamelincourt. Etude et realisation d’une microsonde moleculaire a effet Raman quelques do- maines d’application. PhD thesis. University of Science and Technology of Lille, 1979. GJ Rosasco. Raman microprobe spectroscopy. In: Advances in Infrared and Raman Spectroscopy, Volume 7. London: Heyden, 1980. E Ez. Private communication, 1978, E Etz, Raman microprobe analysis: Principles and applications. Scan Electron Microsc 1:67-92, 1979. DR Clarke, F Adar. Raman microprobe spectroscopy of polyphase ceramics. In: DR Rossington, RA Condrate, RL Synder, eds. Advances in Materials Characterization. Materials Science Research Volume 15. New York: Plenum Press, 1983, pp. 199-214, I De Wolf, Micro-Raman spectroscopy to study local mechanical stress in silicon integrated circuits, a topical review. Semicond Sci Technol 11:139-154, 1996, F Adar, H Noether. Raman microprobe spectra of spin-oriented and drawn filaments of poly(ethylene terephthalate), Polymer 26:1935—1943, 1985, J Barbillat, P Dhamelincourt, M Delhaye, E Da Silva, Raman confocal microprobing, imaging and fibre-optic remote sensing: a further step in molecular analysis. J Raman Spectrosc 25:3-11, 1994. DN Batchelder, C Cheng, W Muller, BJE Smith, Makromol Chem Macromol Symp 46:171, 1991 A Govil, DM Pallister, LH Chen, MD Morris. Appl Spectrosc 45:1604, 1991; A Govil, DM Pallister, MD Morris. Appl Spectrosc 47:75, 1993. T Wilson, ed. Confocal Microscopy. London: Academic Press, 1990. F Adar. Developments in Raman microanalysis. In: RH Geiss, ed. Microbeam Analysis—1981. San Francisco: San Francisco Press, 1981, pp 67~72. 'M Bowden, DJ Gardiner, G Rice. J Raman Spectrosc 21:37, 1990. J Barbillat. Etude d'une seconde generation de microsconde optique a effet Raman mettant a profit les avantages de la detection multicanale. Thesis. Lille, 1983. M Delhaye, E Da Silva, J Barbillat. European Patent 92400141.5 (1992). L Markwort, B Kip, E Da Silva, B Roussel. Raman imaging of heterogeneous polymers: A com- parison of global versus point illumination. Appl Spectrosc 49:1411~1430, 199. SD Schwab, RL McCreery. Versatile, efficient Raman sampling with fiber optics. Anal Chem 56: 2199-2205, 1984, RD McLachlan, GL Jewett, JC Evans. Fiber-optics probe for sensitive Raman analysis, U.S. Patent 4,573,761 (filed September 14, 1983; issued March 4, 1996), PE O'Rourke, RR Livingston. Fiber optic probe having fibers with endfaces formed for improved coupling efficiency and methods using the same. U.S. Patent 5,402,508 (March 28, 1995). P Plaza, NQ Dao, M Jouan, H Fevriet, H Saisse, Simulation et optimisation des capteuts & fibres optiques adjacents. Appl Optics 25:3448-3454, 1986. PJ Hendra, G Ellis, DJ Cutler. Use of optical fibres in Raman spectroscopy. J Raman Spectrosc 19: 413-418, 1988, (a) TF Cooney, HT Skiner, SM Angel. Comparative study of some fiber-optic remote Raman probe designs, Part I: Model for liquids and transparent solids. Appl Spectrosc 50:836-848, 1996. (b) TF Cooney, HT Skinner, SM Angel. Comparative study of some fiber-optic remote Raman probe de- signs. Part II; Tests of single-fiber, lensed, and flat- and bevel-tip multi-fiber probes. Appl Spectrosc 50:849-860, 1996 MM Carrabba, RD Rauh, Apparatus for measuring Raman spectra over optical fibers. U.S. Patent 5,112,127 (filed November 28, 1989; issued May 12, 1992), Copyright © 2001 by Taylor & Franeis Group, LLC 85, 86. 87, H Owen, JM Tedesco, JB Slater. Remote optical measurement probe. U.S. Patent 5,377,004 (filed October 15, 1993; issued December 27, 1994). M Delhaye, E Da Silva, G Martinez. Spectral band filtration spectrometry apparatus, U.S. Patent 5,424,825 (issued June 1995). IR Lewis, PR Griffiths. Raman spectrometry with fiber-optic sampling. Appl Spectrosc 50:12A~ 30A, 1996, SA Asher. UV resonance Raman spectroscopy for analytical, physical and biophysical chemistry. Part 1. Anal Chem 65:59A-66A, 1993; Part 2, Anal Chem 65:201A-210A, 1993. 2001 by Taylor & Francis Group, LLC 3 Raman Spectrometry and Its Adaptation to the Industrial Environment Joseph B. Slater, James M. Tedesco, Ronald C. Fairchild, and lan R. Lewis Kaiser Optical Systems, Inc., Ann Arbor, Michigan |. INTRODUCTION Raman spectroscopy is a valuable tool for studying molecular vibrations, whether in the industrial or the laboratory setting. It can provide high information content via well-sep- arated spectral features in many categories of samples, including liquids, solids, and gases. ‘There are a number of different approaches to designing instrumentation for Raman spectroscopy. This chapter describes the key elements of a Raman spectrometer and the relative merits of different implementations of those elements with respect to their use in the industrial process environment. ll. RAMAN SPECTROMETER SYSTEM OVERVIEW In the last 15 years, there has been a renaissance in Raman spectroscopy due, in large part, to the development of several hardware components, including the following: 1. The commercialization and acceptance of Rayleigh rejection filters as effective replacements for zero-dispersion double monochromators 2. The availability of compact lasers ‘The advent of spectroscopic-grade charge- oupled device (CCD) detectors A refinement in sampling optics and devices (microprobe, fiber-optic probes, and the traditional sample chamber) 5. Availability of powerful personal computers with associated software ae In Chapter 2, the historical impact of these devices have been reviewed. These individual components have been assembled to produce three popular and fundamental types of Raman spectrometers: the laboratory Fourier transform (FT)—Raman system [1,2], the CCD-based microprobe [3], and the fiber-coupled CCD-based process Raman analyzer Copyright © 2001 by Taylor & Franeis Group, LLC [4,5]. These different Raman spectrometers address different questions, as shown in Table 1 The hardware of a Raman spectrometer system may be split into four basic compo- nents: a laser, an optical sampling system, a wavelength separator, and a detector. The laser and spectrometer hardware may be packaged together in a “base unit.” The optical sampling system includes a means for illuminating the sample material with laser light and collecting the Raman scatter for input to the spectrometer. It may be optically inter- faced to the base unit cither through direct imaging or via fiber optics, the latter being particularly preferred in on-line industrial settings. A block diagram of the generic com- ponents making up a Raman spectrometer is given in Fig. 1. The function of the spectrometer is to separate the wavelengths that comprise the polychromatic input Raman signal and present them to the detector. The most familiar class of spectrometers to the traditional Raman spectroscopist is the spectrograph. Spec- trographs are dispersive in nature, using gratings and/or prisms to angularly separate the wavelength components into their individual components prior to measurement at the detector. Another type of spectrometer used for Raman spectroscopy is the Fourier trans- form or FT spectrometer [1,2,6,7]. In Chapter 2, the FT-Raman spectrometer and its impact on the field of Raman spectroscopy has been discussed. Detailed reviews of the theory of operation of this type of spectrometer can be found in several excellent references dedi- cated to FI-Raman spectroscopy by Hendra ct al. [1], Chase and Rabolt [2], and Parker [6]. In an FT-Raman spectrometer, the light to be analyzed is passed through a scanning interferometer, normally a Michelsen interferometer, to generate a temporal signal on a Table 1 Comparison of Different Types of Raman Spectrometer Excitation wavelength Sample ‘Sampling Spectrometer (am) quantity Speed device Market FE-Raman 1047~ Bulk Medium-slow Sample Analytical, routine 1339 chamber High cm" accuracy High concentration Qualitative Quantitative CCD—microprobe — 206.5- Trace —-Fast-medium — Microscope Research lab to process 830 1-4 em™ accuracy Low concentration Qualitative visualization Failure analysis, Homogeneity determination Fiber-coupled 450-830 Trace Fast Fiber-coupled Research lab to process CCD Process probe head 1-2 em accuracy Analyzer Low concentration Qualitative Quanitative Process control 2001 by Taylor & Francis Group, LLC Sampling systems Laser | selection ||>| |&> device device Wavelength separator Detector Figure 1 Block diagram of the generic components making up a Raman spectrometer. single detector. The Fourier transform of that temporal signal corresponds to the wave- length spectrum of the input light. A diagram of a generic FT-Raman spectrometer is shown in Fig. 2. The major drawbacks with the FT-Raman spectrometer approach are as follows: 1. Low sensitivity due to the 1" effect of operating with 1064-nm excitation. Shorter excitation wavelengths have been demonstrated but have not been widely imple- mented [8,9]. 2. Low instrument throughput even for optimized Raman operation using all gold- coated optics (the beam splitter itself imposes a theoretical maximum of 50% throughput) [2]. Improper filling of the collection optics leads to a shift in the observed peak positions [10]. 4, Tnaccuracies in the path of the internal reference (HeNe) laser lead to a linear shift in the observed peak position [10] 5. Mulitplex disadvantage (noise from an intense Rayleigh line or intense Raman peak is distributed over the whole spectrum; in addition, fluorescence or sampling heating, which both can be very intense signals, yield similar sources of noise [1,2). This has lead to FT-Raman spectrometers being limited to observing sam- ples below 300°C, where blackbody emission from the hot sample and associated noise are kept to an acceptable level. It should be noted, however, that consid- Copyright © 2001 by Taylor & Franeis Group, LLC Fixed Mirror HeNe reference laser (632.8 nm) CaF, or Quartz} Beamsplitter, Excitation/ Collection Moving Mirror Lens HeNe ‘l Teeestor| Notch Filters Ge or InGaAs detector \ \Q—__} Figure 2 Diagrammatic representation of @ generic FR: ‘Copyright ©2001 by Tylor & Fas Group, LLC 180° In spectrometer using lens based collection, Sample Rotatable Mirror erable work has been undertaken in the laboratory using pulsed laser excitation to deal with the problem of thermal background [11-16] 6. Due to the multiplex disadvantage, many FT-Raman spectrometers use filters, which make it impossible to observe the Rayleigh line. Thus, any shift in the laser line leads to shifts in the observed wavenumber positions of the sample under investigation. Although an interactive correction using a comparison of the peak positions of Stokes and anti-Stokes, spectral features can be used to identify the current laser wavelength. 7. Sample self-absorption. This phenomenon originates from a competition between Raman scattering and infrared absorption bands [17-19]. In effect, Raman scat- tering is absorbed by infrared absorption, both fundamentals and overtones, before it can escape the sample and be detected. The problem is most severe in the near- infrared (NIR) spectral region, where the first and second overtones of strong CH, NH, and OH are observed A more detailed comparison of the dispersive versus the FT approach is beyond the scope of this chapter. Bowie et al. {20,21] have recently conducted a comparison of these types of instrument. FT-Raman spectrometers involve inherently more delicate moving hardware, which are vibration sensitive, and therefore, they are more suited to the use in the laboratory than in the process environment. Furthermore, many practical applications of Raman spectroscopy involve moving, flowing, turbulent sample materials that in some cases may not be homogencous. Such environments are inherently less compatible with the time-scanned FI-Raman method due to temporal variations in the scattered light in- tensity and/or spectral content being returned from the sample. These sampling consider- ations change the observed scattering parameters and, therefore, can lead to multiplex noise being distributed over the spectrum. In a dispersive imaging spectrograph, each point in the accumulated spectrum represents the same time-averaged sample return; however, there is no equivalent of multiplex noise in the dispersive spectrometer. Gervasio and Pelletier [22] reported a comparison of near-infrared (NIR) dispersive and FT-Raman analyzers for on-line studies of phosphorus trichloride, where the effects of multiplex noise causes by bubbling, particulate movement are discussed. For these reasons, the remainder of this chapter will center on Raman analyzers that use the dis- persive approach. Despite the drawbacks mentioned, there have been a number of reports of the use of FT-Raman spectrometers for on-line analysis (23-32). ‘At this juncture, it is necessary to define what will be considered under the title of process Raman spectroscopy in the industrial environment. Quality assurance/quality con- trol (QA/QC) applications and failure analysis are important areas of industrial interest; however, these areas, in general, can be adequately addressed with standard laboratory Raman spectrometers including FT-Raman spectrometers and dispersive microprobes. For the purpose of the remainder of this chapter, the industrial environment and the process Raman analyzer will be restricted to instrumentation and protocols usable for on-line measurements. At this point, it is necessary to outline a definition of what the requirements for an on-line process Raman analyzer are. A process Raman analyzer should be composed of components which have the following properties: 1. Can be packaged for use in a process environment 2. Require limited utilities 3. Possess a high mean time between failures and little, if any, service Copyright © 2001 by Taylor & Franeis Group, LLC 4, Maintain a high level of calibration over extended periods of time and a variety of environments 5. Use fiber coupling to access the sample material These components must also provide the following: 1. Good sensitivity 2. High resolution 3. Large simultaneous spectral coverage With these criteria in mind, the different instrumental approaches currently available will be evaluated. The remainder of the chapter will be broken down into a preliminary dis- cussion of the components available for dispersive Raman spectroscopy and, subsequently, the unique challenges of the process environment. A. Lasers: An Introduction Due to the very low fraction of scattered photons in a Raman experiment, a high-intensity monochromatic source is required. From the development of the laser [31-37] in the 1960s, the laser quickly became the only practical excitation source for Raman spectros- copy. There are many well-established lasers for use for Raman spectroscopy which are completely unsuitable for process Raman spectroscopy, including the titanium-sapphire solid-state laser and the dye laser. Of the suitable laser types, there are several basic requirements which must be met for fiber-coupled process Raman spectroscopy; these are wavelength stability (around 0.05 nm or less), high spatial mode quality (so that a large percentage of the light can be imaged into standard fibers), and moderately narrow line- width (0.05 nm or less). The absolute values required ultimately depends on the applica- tion, but the large majority of process applications share similar needs. Initially, multiwatt lasers were necessary to produce a Raman spectrum in the laboratory. Currently, laser power in the tens to hundreds of milliwatts are adequate to provide the necessary sensitivity for most process measurements. In the late 1960s and early 1970s, powerful mainframe ion gas lasers such as argon or krypton [38] were used as Raman excitation sources. Many laboratory systems still employ them, and they easily meet the basic performance needs of laboratory Raman systems, These mainframe lasers, however, are not well suited for implementation in pro- cess Raman systems because they tend to be physically large (typically over | m in length for the laser head and a separate power supply), require enormous amounts of power (often several kilowatts), require three-phase power, require cooling water, require filtering to remove plama emissions, and are relatively unreliable (a new laser tube is required every 13000 hr). Over the last 15 years, air-cooled compact versions of these lasers have become available. These lasers still suffer from a separate head and power supply structure, they require filtering to remove plasma emission, and they are still relatively unreliable (tube change every 3000-5000 hr). Generally, air-cooled ion lasers are limited to single- wavelength output powers of less than 200 mW in the most intense laser line. Finally, these lasers produce an enormous amount of heat, which is normally dispersed using cooling fans. In a laboratory environment, heat dissipation to the air-conditioned laboratory air is generally trivial. In a process environment where the laser may need to be enclosed into a process enclosure, heat dissipation may be difficult, and, in general, ion lasers are unsuitable, In the QA/QC process environment such as the computer hard-drive industry Copyright © 2001 by Taylor & Francis Group, LLC (Chapter 24), where the Raman system is installed in a laboratory environment, these lasers are still popular choices. In a more general sense, it is the advent of diode lasers and diode-pumped lasers that has opened the door to process Raman spectroscopy. As of this writing, the list of available lasers suitable and tested for process Raman includes frequency-doubled diode-pumped Nd:YAG at 532 nm, HeNe at 632.8 nm, and external cavity stabilized diodes ranging from 780 to about 850 nm. 1. Helium—Neon Ion Laser The gas laser most suitable for use in the process environment is the helium—neon (HeNe) laser [39]. The HeNe lasers are extremely mature products, being produced by the millions over several decades for thousands of applications. As a result of the production volume, these lasers are relatively inexpensive, with a cost from a few hundred to several thousand dollars for 1-30 mW of power. In addition to the attractive price of these devices, the reliability is also excellent, with rated lifetimes in the 100,000-hr range. The main li tations with HeNe lasers is power per unit length and total output power. The physics of the HeNe laser places a limitation on how much power can be generated for a given length of laser cavity. Thus, a 30-mW HeNe laser (of which only about 20 mW may be of sufficient mode quality to be usable for fiber coupling) is about 1 m in length. At double the cavity length, laser powers of 40-50 mW are achievable. This size~power relationship makes process packaging a practical HeNe-based Raman system difficult. 2. Diode-Pumped Solid-State Lasers Solid-state crystalline materials such as a Nd: YAG (neodymium:ytrium aluminum garnet) can yield lasing action at 1064 nm when pumped with NIR (809 nm) radiation (36,40). Initially, multiple high-powered flash lamps were used to pump the crystal. These flash lamps required an enormous amount of electrical power to operate and, in addition, gen- erated a significant amount of heat. The heat was typically conducted away using cooling water, In addition, the flash lamps were a major reliability concern requiring changing every 500 hr. In the early 1990s, diodes were identified which could be used to pump solid-state lasers. In these lasers, a diode that would not meet the performance requirements for a laser suitable for Raman spectroscopy is used to pump another laser media. The solid-state laser being pumped is more stable and narrow, with better mode quality than the diode laser itself. In the case of the diode-pumped Nd: YAG laser, diodes lasing in the 809-nm range are used to pump a Nd:YAG crystal, which then lases at 1064 nm. A further increase in wavelength flexibility is provided by frequency doubling. Certain crystalline materials such as KTP (potassium trihydrogen phosphate) or KDP (potassium dihydrogen phosphate) exhibit optical nonlinearity; the optical properties of the material change with the intensity of the light passing through it. Without going into the details, as this information is readily available elsewhere [36], the nonlinearity allows the crystal to convert two photons into one photon at half the original wavelength (i.e., twice the original frequency, 532 nm for a 1064 nm Nd:YAG laser). The doubling process for a diode-pumped Nd: YAG laser is energy wasteful; conversion is typical between 5% [36] and 18% [40]. The result of using diode pumping in solid-state lasers is that these lasers offer longer lifetimes (5000 hr) than their ion laser (514.5 nm) equivalent (2000 hr), are small and compact, only require 110 V power, produce minimal heat, and can be operated up to multiple watts of laser output without cooling water. The disadvantage to diode- pumped solid-state lasers is their expense. Copyright © 2001 by Taylor & Franeis Group, LLC 3. Diode Lasers An alternative strategy is to use the laser diodes [36,41] directly, and thus more efficiently, in an external-cavity configuration as described in many excellent references [42-44]. Briefly, these lasers work by exciting a gain material sandwiched between two reflectors (the cavity). The excited material produces photons that reflect back and forth between the mirrors, thus resonating in transverse and longitudinal modes supported by the gain medium, reflector spacing, and waveguide geometry. This resonant effect is what produces the intense and pure light associated with lasers. The difficulty in producing laser diodes that meet the stability and mode quality requirements has to do with the fact that the diodes have a very tiny cavity and the lasing material has a very broad gain characteristic. These factors cause the diode to lase in multiple modes, either simultaneously (in multi- mode structures) or “hopping” in random sequential fashion between a number of wave- lengths as photons of different wavelengths and directions compete for dominance (in a single-mode structure). Furthermore, the lasing wavelength is also very temperature sen- sitive; thus, there are stability problems with free-running laser diodes. This problem can be reduced by placing the diode in a temperature-controlled environment, but to truly stabilize the laser output, a wavelength-selective external cavity is required. In most widely adopted external cavity approaches, the diode is placed inside a larger (external) cavity designed to narrow the selection criteria for excited photons. This is usually accomplished by replacing one of the diode facet reflectors with an external re- flector containing a diffraction grating. The purpose of the grating is to severely restrict, the range of wavelengths that can resonate with the single remaining diode reflector. Thus, the external cavity forces the diode to operate in a narrow, stable emission mode. A final consideration for these devices is filtering to remove superradiance. This superradiance appears as wings on either side of the laser line, Both dielectric and holographic filtering approaches have been used to remove this superradiance; however, holographic-based devices offers high suppression of the superradiance combined with very high throughput of the laser line (stimulated emission). Although this technology is relatively new to com- mercial application, it has already made a significant impact on the field of process Raman spectroscopy where both 780-785- and 830-nm lasers have been adopted [45—50]. Most recently, diode lasers lasing at 670 nm have been introduced which offer approximately 100 mW of laser power in a small compact unit [51]. These lasers may, in the future, become a replacement for or an attractive alternative to the HeNe laser. Current commercial stabilized diode lasers are small in size per unit power (less than 1 ft Jong and occupying less than 0.1 ft° for several hundred milliwatts), consume modest amounts of power (less than 50 W), and can be durable. 4. Laser Reliability Lasers, at least those employing diodes, generally set the reliability limit for process Ra- man systems. Whereas HeNe lasers can continue to operate for more than a decade in continuous use (90,000 hr), the current generation of laser diodes suitable for Raman usually possess mean times between failures (MTBF) between 9000 and 18,000 hr. Al- though this lifetime is excellent for a laboratory setting when compared with traditional laboratory lasers (e.g., argon ion, 2000 hr), a MTBF of 45,000 hr (5 years) are closer to what the process industry would like. In the case of diode lasers, technical advances, driven primarily by the telecommunication market, have led to new diode materials which in the next few years will enter the commercial mainstream. These materials are anticipated to yield lifetimes equivalent to the HeNe laser. 2001 by Taylor & Francis Group, LLC At this point, there are a few choices the system designer can make that can signifi- cantly improve overall system reliability. The brute-force approach is to use a heavily oversized diode and turn the power down. Diode lifetimes are considerably longer when run at a reduced current. This approach, however, yields limited potential results in fiber- coupled Raman spectroscopy because higher-power-laser diode apertures are larger and therefore do not couple efficiently to small-core optical fibers. Diodes operated at lower temperatures also last longer. Another approach possible for a process Raman analyzer is to design the analyzer with two lasers: when one fails, the failure is detected and the backup is automatically activated while the operator is notified. The failed laser may then be replaced off-line by a technician at a convenient time. Despite the fact that this approach increases the initial capital cost of the analyzer, it has been fielded for critical installations where continuous operation is necessary. B. Laser-Line Rejection A unique requirement of Raman systems that further differentiates them from more tra- ditional spectrographic analyzers is the requirement for filtering out the radiation at the laser wavelength. The laser-line rejection device filters out both reflected laser radiation and Rayleigh scatter, leaving only the much weaker Raman scattering signal prior to the detector. There are two types of filters used for Raman spectroscopy: the edge filter and the notch filter, These filters differ in that for a notch filter, a range of wavelengths are blocked, but information at higher and lower frequencies can be observed. In the edge filter, typically all wavelengths above or below a certain wavelength are blocked. Without laser-line filtering, the scattered radiation at the laser wavelength will be approximately eight or more orders of magnitude more intense than the Raman signal of interest, making the Raman signal virtually impossible to detect (due to the limited dynamic range of the detector). The ideal notch filter would reduce the intensity of a small wavelength range containing the laser line by eight-plus orders of magnitude while transmitting 100% of all other wavelengths. The following subsections contain a brief survey of the most popular notch-filtering technologies: zero-dispersion monochromators, absorption filters, and re flectance interference filters. There are, however, other filtering approaches that have been proposed for Raman spectroscopy, such as the colloidal filter [52], the semiconductor band- gap filter [53,54], and the iodine vapor filter [55] to name a few. 1. Zero-Dispersion Double Monochromator An early approach for creating a notch filter was implemented in the triple monochromator shown schematically in Fig. 3. In Chapter 2, the development and capabilities of the triple- monochromator-based spectrometer are discussed, and thus only brief mention will be made here. In this approach, the composite Rayleigh/Raman signal is fed into a spectro- graph, usually a Czerny—Turner. At the output image plane of this first spectrograph stage, the laser and Raman signal are dispersed and, hence, spatially separated. A mask of some sort, either a wire or a knife-edged blade, is used to block the laser (or the laser and all the wavelengths below, in the case of the blade). The radiation with the laser removed is then fed into a second spectrograph that is identical to the first, but arranged with mirror image symmetry. This has the effect of undoing the dispersion of the first spectrograph (this configuration is often referred to as a zero-dispersion spectrograph). The resulting undispersed Raman signal is subsequently redispersed by the third spectrograph (Fig. 3) It should be noted that one of the key advantages of this spectrometer is that by increasing the linear dispersion of the filtering spectrograph stages, it is possible to filter to very close Copyright © 2001 by Taylor & Franeis Group, LLC Single stage spectrograph Ms Detector | Jus we M3) { Jo si Zero-dispersion double monochromator Figure 3 Generic schematic of a triple monochromator featuring a zero-dispersion double-monochromotor Taser-ine filter. Copyright ©2001 by Taylor & Francis Group, LLC to the laser wavelength, Commercial triple spectrographs can attenuate the radiation at the laser wavelength by a factor up to 10" and allow signals as close as a 5-cm’' Raman shift to be detected. A further advantage of this approach is that the notch is continuously tunable over a very large spectral range. From its introduction in the early 1980s until the early 1990s, this method of notch filtering and this spectrometer were established as the preferred configuration for laboratory-based dispersive Raman spectroscopy. The disad- vantages of this approach, however, are that the spectrometers are physically large, me- chanically complex (leading to multiple potential sources of failure and calibration errors), costly, and have poor throughput due to all of the spectrograph stages and associated optics. For these reasons, this method of rejection of the radiation at the laser wavelength is unsuitable for process Raman spectroscopy. 2. Absorption Filters This type of filter can be dated back to Raman’s original experiments [56]. During these ground-breaking experiments, Raman used colored-glass absorption filters to observe the shifted radiation and demonstrate the effect that now bears his name [56]. In general, however, filters based on dyes have absorption bands that are much too broad to be of use in modern Raman spectrometer systems. Atomic vapor-absorption bands can be used to produce a high-performance notch filter [57-65]. As its name suggests, this type of filter relies on the absorption of specific wavelengths by electron orbital transitions in an atomic vapor. The filter is based on a metal vapor (¢.g., rubidium [65]), which is often held in a transparent glass cell. Such filters can produce fair attenuation (3-5 optical density) over a very narrow bandwidth (less than 1 cm“) [63]. Their principal drawbacks are their relative complexity, their many closely spaced absorption peaks that can interfere with the data of interest, and the fact that the absorption peaks are not necessarily coincident with common laser wavelengths. Because the filter's absorption bands are so extremely narrow, if the laser shifts frequency even minutely, it will fall outside the absorption peak and will not be blocked. These challenges reduce the general applicability and attractiveness of this type of filter for Raman systems. 3. Reflective Interference Filters The adoption of reflective-interference filters in the late 1980s has undoubtedly revolu- tionized the field and has made a significant contribution to making process Raman spec- troscopy possible, The reflective-interference filter is currently the preferred choice for laser-line removal for many applications in both the research and the process Raman field. This type of filter operates as a wavelength-selective mirror, reflecting away a narrow band while transmitting all others. These filters operate by producing a series of partial reflec- tions through their thickness (Fig. 4(a)). The spacing of these reflections is such that for a given wavelength, incident at a specific angle, all the reflections from periodic boundaries constructively interfere [66]. Thus, the given wavelength at the specified angle is reflected with high efficiency. Other wavelength and angle combinations do not meet the construc: tive-interference criteria and pass through unaffected. There are two basic types of reflec- tive-interference filters: dielectric and holographic. The principle drawback to these filters is their wavelength specificity. This initially limited their acceptance in the laboratory environment, but due to the specificity of each process Raman application, the limited number of laser wavelengths available, the great simplicity offered by using a filter instead of multiple monochromators, and the need for high optical throughput, filters were quickly adopted [4,67]. Copyright © 2001 by Taylor & Franeis Group, LLC 1 ® 15 frown © 0 16 30 45 60 Refractive Index Film Thickness, microns Figure 4 Schematic representation of filter thickness, bulk refractive index, and index modulation for dielectric, holographic edge, and holographic notch. (From Ref. 84; reproduced with permission.) a. Discrete Multilayer Dielectric Stack Filters. Multilayer dielectric filters are con structed using standard vacuum-deposition techniques with inorganic materials such as titanium dioxide (TiO,), silicon dioxide (SiO.), and tantalum oxide (Ta,O,). Multiple layers of dielectric, transparent glasslike materials are built up in succession on a glass substrate. The simplest approach is to use alternating layers of two different materials having a slightly different refractive index. Each layer has an optical thickness of one-quarter wave. The peak OD is determined by the total thickness and the refractive index differential. The reflective bandwidth is proportional mainly to just the refractive index differential. Dielectric interference filters can be made with hundreds of layers to reflect either a very narrow range of wavelengths at a reasonably high OD, up to about OD 4 (acting as a notch filter), or be constructed in a different way to produce a long-pass filter. In the case of the long-pass filter used for Raman spectroscopy, only the Stokes—Raman-shifted ra- diation can be observed. Both dielectric notch and long-pass filters are commercially avail- able from a number of manufacturers [68,69], are compact, can be relatively inexpensive (depending on the performance level), and are durable. This filters can be manufactured to operate from the deep ultraviolet (UV) to the NIR and have been employed in Raman systems varying from dispersive UV-Raman [70-72], to visible Raman [73-75], to FT- Raman [76] One of the principal disadvantages of dielectric notch filters is “filter ringing.” This is inherent to the fundamental physics of discrete-layer interference filters [66]. The abrupt changes in index between the successive layers of a dielectric filter causes ringing artifacts in its reflective characteristics (Fig. 5). Instead of reflecting just the wavelength range of interest, it also weakly reflects other wavelength ranges throughout the Raman spectrum of interest. The spectral locations of these artifacts are sensitive to temperature. Thus, the artifacts, in addition to reducing system throughput in some regions, also make it more difficult to calibrate the instrument intensity axis. In some extreme cases, the filter ringing, can become indistinguishable from the Raman data. A more preferred refractive index profile is the sinusoid. Theoretically, this profile produces a filter with very little artifact. A relatively new type of dielectric filter, called a Copyright © 2001 by Taylor & Francis Group, LLC 100 % 450 490 530 570 600 Wavelength, nm Figure 5 Plot of transmission versus wavelength for dielectric and holographic notch filters. (Re~ produced with permission from Kaiser Optical Systems, Inc.) rugate filter, seeks to reduce ringing artifacts by approximating a sinusoidal refractive index profile [66,68,77,78]. Rugate filters are also made by vacuum deposition, but instead of laying down discrete layers of individual materials, the materials are deposited simulta- neously in varying proportions, In this way, the index can be made to change gradually up and down as the deposition proceeds. The manufacture of rugate filters is relatively new, difficult, and expensive. As of this writing, only one or two companies are moving this technology into the commercial arena [68] For dielectric rejection filters, the low Raman shift cutoff lies somewhere between 200 and 600 cm”', making observation of many inorganic materials difficult. Lewis et al. [79] were able to extend the low-wave-number range of these filter by demonstrating that by angle tuning the filters, the lower-wave-number cutoff can be shifted to a lower wave number, albeit at the expense of laser attenuation ability. The rejection efficiency of a dielectric filter can be improved upon if multiple dielectric filters are used. There are several physical arrangements of the multiple dielectric filters including what has been termed the multistage filter by Hirschfeld and Chase [76], but the most popular configuration is the “chevron” configuration. Hirschfeld and Chase [76] reported the use of this type of filter in their seminal work describing the development of FT-Raman spectroscopy. The characteristics of the filter used [76] “were a 70% passband filter over the Raman frequency band with a 10” rejection notch at the laser line fre- quency.” Puppels et al. [80,81] and Futamata [75] have also reported the use of the “‘chev- ron” arrangement with different numbers of reflections for Raman spectroscopy in the visible spectral region. Puppels et al. [81] reported a notch of 10°* and a 60-70% trans- Copyright © 2001 by Taylor & Franeis Group, LLC mission of the Raman signal up to 2000 cm”', whereas Futamata [81] reported a notch of 10~* with more than 90% transmission of the Raman signal from 50 to 2000 cm b. Holographic Filters. ‘The most popular current choice for removing laser light for Raman spectroscopy from 324 to 1339 nm is the holographic notch filter (67,82-90). Holographic filters are produced by exposing a photosensitive material, currently dichro- mated gelatin (DCG), to an interference pattern formed by two laser beams [66]. The interference fringes are an intensity pattern whose profile is sinusoidal. This pattern is recorded in the DCG as a sinusoidal refractive index profile. The filter is packaged between two layers of glass in order to protect the DCG layer. Commercial holographic filters can reduce the laser signal by four to eight orders of magnitude while passing as much as 90% of the out-of-band signal (depending on the filter type used, Notch or Super-Notch 184,85,91]). They provide acceptable transmission to Raman shifts as low as 50 em‘, Out-of-band reflection artifacts are typically at least an order of magnitude less than those found in dielectric filter Despite their widespread acceptance, holographic filters do have some drawbacks. The DCG material is sensitive to water and, thus, the filters must be completely sealed against moisture. In addition, these filters do not handle temperatures much above 80°C for long periods. This temperature limit is high enough for most Raman applications. In addition, due to absorption in the DCG film, these types of filters cannot be manufactured for use for UV-Raman spectroscopy [92,93]. However, Kim et al. [88] have demonstrated that holographic notch filters can be used with a krypton ion laser operating at 324 nm, Finally, these filters can produce a weak background, which is due to fluorescence generation in the DCG film, which may be seen when observing a very weak Raman band. Lewis et al. [79] noted that dielectric filters can be angle tuned to shift the low-wave- number edge of the filter to a lower wavenumber. In a similar way, holographic-edge [83] and notch filters can also be angle tuned [94]. In addition, these filters can be arranged into a high-reflection-angle multifilter configuration to provide improved low-wavenumber performance [95] C. Spectrograph The basic components of a dispersive spectrograph are a slit or input aperture, an input collimator, a grating, and an output collimator. Light enters the spectrograph through a slit or input aperture, which serves to spatially localize the light. The input collimator collimates the light from the slit and directs it to the grating where it is dispersed, or angularly separated, into its constituent wavelengths. The output collimator images the light from the grating onto the detector. In effect, an image of the slit is being “relayed” by the collimators to the detector. The grating produces a different laterally displaced image of the slit at each wavelength of the input light. The term spectrograph implies a two-dimensional, or “imaging,” device. The sepa rated wavelengths of the spectrum are imaged simultaneously along a horizontal axis at the output image plane. Similarly the vertical axis can represent a multitude of individual inputs, from different probe heads [67] or from different spatially resolved areas of a line- focus microscope [96]. This is distinct from a monochromator, which is inherently a one- dimensional or single-wavelength device [97-99]. During the period 1965-1980, scanning monochromators were the tool of choice. This was primarily due to the fact that the only suitable detectors available were photomultiplier tubes, which have only a single detecting element. The spectrum was scanned across the detector one wavelength at a time, usually Copyright © 2001 by Taylor & Francis Group, LLC by rotating the grating, and the spectrum was built up accordingly. The advent of multi- channel detectors, initially photodiode arrays and, subsequently, charge-coupled device arrays, has largely rendered the scanning monochromator obsolete for the majority of Raman applications [3,100—103]. However, the technique of scanning is still used in some hybrid imaging/scanning systems [104—108]. 1. Spectrograph Design Issues As noted earlier, there are many different types of spectrograph that can be used for Raman spectroscopy in a research laboratory [2,94,108,109]. The method of laser-line filtering, the required spectral resolution, the wavelength range or ranges of interest, the stray-light performance of the spectrometer, and so forth can determine which type of spectrometer is most appropriate for a particular application. As noted previously, the criteria of the process environment impose some unique requirements on a spectrograph. ‘The spectrograph is the heart of the Raman system and has been the focus of much development and change over the past decade. Although an in-depth study of spectrographs and their design is beyond the scope of this chapter, some fundamentals are needed to evaluate the trade-off between the many different approaches used in modern Raman spectrometers. Both one-dimensional spectrographs, those that produce a single strip of diffracted light at the output plane, and two-dimensional spectrographs, those that produce two or more strips of diffracted light at the output plane, will be discussed. a. Gratings. There are two basic types of grating: reflection and transmission [110— 114]. Reflection types have a surface relief grating constructed on their active surface. The relief pattern can be either mechanically ruled [classically ruled diffraction grating (CRDG)], holographically formed [holographically recorded diffraction grating (HRDG)], or replicated from either type of master. As of the time of this writing, HRDGs have completely replaced CRDGs in new commercial Raman instruments. The reflection grating can be described as possessing a series of small stepped or grooved mirror elements, arranged with appropriate spacing and depth such that for a given incidence angle, the output reflections constructively interfere at a different angle for each wavelength. The net effect is to reflect each wavelength at a different angle. This process of efficiently bending light through the use of periodic phase change is called diffraction. Diffractive structures inherently produce an angular separation of different wavelengths, which is called dispersion. Surface-relief transmission gratings operate similarly to reflection gratings, except that the surface relief produces a stepped-phase discontinuity in transmission through a medium of higher refractive index, as opposed to a reflective-phase discontinuity in air. As such, a surface-relief transmission grating requires about four times the physical depth of a surface-relief reflection grating in order to achieve a similar phase modulation and effi- ciency. This greater depth imposes more problematic practical fabrication and performance issues. Scattering from the deeper-groove pattern as well as from inhomogeneities in the refractive medium lead to higher stray-light levels and greater susceptibility to damage and contamination, Therefore, surface-relief transmission gratings are a rarity. There are few, if any, commercially available surface-relief transmission gratings suitable for a Ra~ man spectrograph at the time of this writing. ‘A newer transmission-grating technology is called volume-phase holography, which has many attractive properties to the designer of a Raman spectrograph [109,115,116]. Volume-phase gratings operate by forming refractive index variations in the volume of an optically thick (several wavelengths of light) material. These refractive index variations Copyright © 2001 by Taylor & Franeis Group, LLC form a continuous-phase diffractive structure that can yield a very high dispersion and a very high diffraction efficiency, as predicted by coupled-wave theory [117]. Typically, this type of grating is manufactured using a holographic process or the interference of two laser beams in a photosensitive film of dichromated gelatin or photopolymer [67,118]. The resulting grating, having no relief structure, may, therefore, be sealed between two pieces of glass. This yields a very sturdy, cleanable element that can also exhibit very low scat- tering, particularly if formed in dichromated gelatin. Further references to transmission gratings will assume the volume-phase holographic grating. There are advantages and disadvantages to both reflection and transmission gratings. The reflective grating’s primary strengths are in having a broad wavelength range and the ability to scan. Disadvantages include efficiency artifacts and susceptibility to contami- nation, Reflection gratings suffer anomalous “dropouts” (Woods anomalies (112]). These are large changes in diffraction efficiency occurring over a fairly localized spectral range, making it more difficult to calibrate a system’s intensity response. This effect can be highly polarization dependent. Because the grating is a surface-relief pattern open to the envi- ronment whose features are on the micron scale, contamination is a major concern. Dust, fingerprints, and so forth can easily degrade the grating. The fragile nature of the surface makes cleaning a reflection grating virtually impossible. The final disadvantage of this type of grating is related to its scanning ability. This ability is offset by the fact that the wavelength calibration of a spectrograph based on a reflective grating is very sensitive to minute changes to its angle, A change in the grating angle of ¢ will cause a 2¢ change in the dispersed output. This places a burden on the mechanical stability of the reflective- grating spectrograph, which can become onerous in the process environment. The relative advantages and disadvantages of volume-phase transmission gratings are quite complementary to those of reflection gratings. Transmission gratings are completely sealed in glass and thereby virtually immune to contamination, If the exterior glass does get dirty, it can be easily cleaned without causing damage. This robust construction is perhaps the greatest attraction of the volume-phase grating in the process environment. Furthermore, volume-phase gratings have a smooth efficiency profile with no anomalous artifacts. They are over an order of magnitude less sensitive to rotations than either CRDGs or HRDGs. This, of course, means that they cannot be angle scanned to extend the wave- length range to any useful degree. They diffract efficiently over a relatively narrow range of wavelengths in comparison to reflection gratings. However, their efficiency spectral bandwidth is wide enough for most Raman applications (0-4400 cm”! for 532-nm exci- tation). This is why volume-phase gratings have been used in both process and laboratory Raman applications, despite the fact that they are not normally found in general-purpose (e.g., UV/vis) spectrographs. A related advantage of the volume-phase transmission grating is that because the grating is not moved during a spectral acquisition, the task of spec- trometer calibration is made easier. b. Throughput: Etendue/Vignetting. The amount of light that an optical system is capable of collecting is referred to as its etendue, which is the product of the area of the collection aperture and the solid angle of the light bundle that can be accepted and detected through that aperture. Etendue is a global parameter, where the smallest value anywhere in the optical path is the etendue of the overall system. In the design of fiber-coupled Raman systems, the engineer typically has the least amount of control over the fiber optics (assuming the use of standard fibers). Therefore, the etendue of the spectrograph should be equal to or greater than that of the fiber coupled to it. The most common multimode Copyright © 2001 by Taylor & Francis Group, LLC fibers have apertures of between 50 and 200 jum, with acceptance cones of around ff2. In a technique called “‘/-number matching,” optics are used to produce a magnified image of the collection aperture (fiber in this case) which reduces the ff# by the same measure. ‘Thus, an f/2 fiber can be efficiently coupled to an /i4 spectrograph. The penalty is that the magnified image will degrade the spectrograph’s resolution, range, or both Another factor related to throughput is vignetting, which is simply the reduction of optical aperture at off-axis field angles, demonstrated in Fig. 6. In spectrograph design, this truncation is primarily confined (o the output collimator. If the beam aperture of the input collimator is 1 in., it is easy to see that an output collimator with an aperture of 1 in, will collect all the light. This is true, however, for only the single wavelength that illuminates the collimator on-axis. Other wavelengths will have their -in. apertures exiting the grating at different angles and will be increasingly vignetted the farther away they are from the “center” wavelength. Thus, the output collimator needs a larger aperture than the input collimator to pass all the light to the extremes of the wavelength range. Many spectrographs are quoted only by the //# at this “center” despite the fact that it can be significantly reduced toward the edges of the spectral field. It is apparent that the closer the output collimator is to the grating, the less vignetting will occur for a given /## output collimator. Spectrographs based on transmission gratings have a significant advantage of low vignetting relative to spectrographs that use reflection gratings. This is because the input and output collimators fall on opposite sides of a transmission grating, allowing them to be physically much closer to the grating, thereby reducing the vignetting at the output aperture due to off-axis angles. ©. Range and Resolution. In discussing the simultaneous spectral range and reso- lution of a spectrograph, a number of factors are brought together. These include the grating, the collimators, the slit, and the detector. For any given set of parameters, range and resolution are traded off against each other. Collimated Beam from grating 4— Vignetted light ZZ Collimator LEN mirror Center wavelength Detector plavie Figure 6 Principles of vignetting; example for a reflective-based collimator. Copyright © 2001 by Taylor & Franeis Group, LLC The simultaneous spectral range is practically defined as the difference between the lowest and the highest wavelengths that just fit on opposite edges of the detector array. The main parameters that affect the simultaneous spectral range are the focal length of the collimators, the amount of dispersion of the grating, and the overall size of the detector. These parameters can be simplified somewhat by combining the collimator focal length and grating dispersion into a single factor called linear dispersion, which is simply the amount of wavelength dispersion per unit length at the detector. In scanning spectrographs, one can select parameters to have a high resolution with a limited simultaneous spectral range and then scan the full range of interest across the detector. The disadvantages to this in the process environment are the mechanical com- plexity and stability concerns with the scanning hardware, plus the fact that Raman signal photons outside the instantaneous range of the detector are being lost. Furthermore, scan ning spectrographs suffer the same disadvantages with respect to moving or turbulent media as the scanning interferometer FT systems. A more efficient and universally appli- cable solution is to acquire the entire spectral range of interest simultaneously. The advantage of imaging spectrometers that collect the entire spectral range of in- terest simultaneously on a two-dimension detector when compared to scanning mono- chromators are the following: No moving parts Complete spectral coverage (which obviates problems when the background level or signal at only a few bands changes) 3. A large reduction in time when collecting a complete spectrum 4. Small instrument footprint Resolution can be a complex and misunderstood subject, and the reader is encouraged to look at Ref. 119 for more detail. For this discussion, resolution is the closest separation between two monochromatic wavelengths that can be resolved [119]. The factors affecting resolution are linear dispersion, the size of the slit or input aperture, the image quality of the optics, and the size of the individual detector elements. Once again, a series of trade- offs can be made. Increasing linear dispersion improves resolution but reduces range. Increasing the size of the input aperture (or fiber diameter in the case of a fiber-coupled process Raman analyzer) increases the etendue but reduces resolution For some of the factors, more is simply better, but often difficult to realize in practice. Better image quality improves resolution without affecting range, but usually requires more complex imaging optics, More detector elements can improve resolution, but only if the input aperture is correspondingly reduced in size. d. Stray Light. The stray-light ratio is a particularly important factor in the design of spectrographs used for Raman data collection. Stray light is caused by photons being scattered from their intended paths. The scattered photons create a background level that will interfere with the weak Raman signal. It is usually expressed in optical density (OD), which is, to say, a ratio of scattered light to incident light in factors of 10. For example, OD 4 is an attenuation of 10~* Commonly available reflection gratings, particularly CRDG types, tend to have slightly higher stray light than transmission gratings. HRDGs, generally, have a better stray-light performance than CRDG [TY Handbook (1980)]. This is due to several factors, such as the large number of relief edges on the surface and the fact they are more sus- ceptible to contamination, 2001 by Taylor & Francis Group, LLC The stray-light performance of reflective and transmissive collimators is nearly equiv- alent, Scratches and dust tend to scatter more light through higher angles on reflective surfaces, but there are generally many more surfaces in transmissive optics. There is also scattering from the bulk material in the transmissive case, which is not an issue in reflective optics. Triple-stage spectrographs produced the ultimate in stray-light performance, achieving, 12 OD. These designs are based on an all reflective optics (Fig. 3). The disadvantages of this design, in both the laboratory and the process environment are (1) low throughput, (2) mechanical complexity, (3) reliability, including calibration stability, and (4) expense. These stray-light levels were necessary when one of the primary functions of the spectrometer was to reject the laser line. The use of laser-line filters to attenuate the laser line by six to eight orders of magnitude leads to a relaxation of the requirement for stray- light performance of the spectrometer. Stray-light values of 4—5 OD are generally good cnough for most Raman measurements and are obtainable with single-stage spectrographs. These can be manufactured in either transmissive or reflective varieties. e. Imaging Performance, Point-Spread Function. The imaging performance of the collimators, as mentioned in Section II.C.1.c, can have a significant impact on the reso- lution of the spectrograph. The extent of the impact will depend on other spectrograph parameters such as size of the input aperture, the focal length of the collimators, and size of the exit aperture (in the case of a single-point detector) or detector elements (in the case of a one- or two-dimensional array). The maximum number and signal separation of multiple inputs is directly affected by spectrograph imaging performance. Although there are a many ways of describing the imaging performance of collimators, the point-spread function (PSF) is perhaps the most useful for spectrographs. An ideal collimator would produce an infinitesimal spot at its focal plane, given a perfectly colli- mated input. A rigorous treatment of PSFs can be found in many optical references [120] For our purposes, the PSF can be thought of as the amount of blurring or “spread” of this spot generated by the actual collimator. In reality, the wave nature of light places a lower limit on this spot size for even a perfect optic. This is known as the diffraction limit and is dependent on the f/# of the collimator and the wavelength of the light. The diameter D of a diffraction-limited spot is approximated by the formula D = 1.22A¢/i#), where A is the wavelength, and j# is the f-number of the collimator, thus giving a minimum achievable spot size range of about 2—5 zm for typical f14 spectrometers operating in the visible to NIR wavelength ranges. Diffraction-limited performance over an extended spectral field is difficult to achieve in practice, being more problematic the lower the //# of the collimator. Fortunately, it is also a overkill for most practical spectrometers. Input apertures are typically 25 um to several hundred microns, with detector element sizes ranging from 5 to 50 ym. PSFs significantly smaller than the lowest of these values will have limited impact on spectral resolution or the number and separation of multiple inputs. ‘An aspect of spectrograph imaging performance that is often overlooked is the off- axis PSF. It is relatively straightforward to achieve a PSF within a factor of 10 of the diffraction limit at one point on the detector. What is much more difficult is to maintain this performance over the full detector width, known as the image field. Many spectrograph manufacturers will specify image performance at the best point and ignore the fact that the PSF may be more than an order of magnitude larger at the edge a 1-in.-wide field. ‘Some spectrographs employing refractive multielement collimators have less than a 50% variation across the same 1-in. field. Copyright © 2001 by Taylor & Franeis Group, LLC Of final note for imaging performance is chromatic abberation. Traditional spectro- graph designs employing reflective collimators are not affected by changes in wavelength. ‘This is not true of the refractive multielement collimators just mentioned. The PSF will have a dependence on wavelength that can be minimized, but not eliminated. Chromatic aberration is generally not a problem if the spectrograph is to be used with a single excitation source, thus limiting the range of wavelengths to within 100 nm or so, as is typical of process Raman analyzers. However, this issue can be a major factor if the spectrograph is to be used with multiple sources common in research applications. 2. Spectrograph Configurations a. Czerny—Turner. The most widely used spectrograph configuration for Raman spectroscopy is the Czerny—Turner [121,122] and its variants (shown in Fig, 7). Com- mercial versions of the Czerny—Turner spectrograph are available from JY Horiba [123], Chromex [124], and Acton Research [125], among others. The Czerny—Turner makes use of mirrors as collimators in an off-axis configuration and employs a planar reflective grating in the collimated space. The Czerny—Turner design leads to a basic one-dimen- sional (1D) spectral dispersion dimension. The off-axis configuration generates significant optical aberration, most notably astigmatism. It is for this reason that most modern Czerny—Turner spectrographs use aspheric mirrors. This allows for the correction of the astigmatism at a small area at the center of the image plane. The correction degrades rapidly away from the center of the field. The presence of these aberrations places a practical lower limit on the //# of a Czerny-Turmer spectrograph as aberrations increase exponentially with decreasing fi#. For a Czerny—Turner operating, at /74, the spot size at the edge of a modern wide-format charge-coupled detector (CCD) is typically 140 um for a 25-ym input. Because most fiber optics used in process Raman systems operate at around {J?2. or lower, this makes the Czerny-Turner less than ideal as an imaging spectrograph for these applications. Spectral resolution can be recovered to some extent by increasing the Mirror 2 Single Grating or Multi-Grating Turret ( Mirror 1 | | Entrance | Slit Figure 7 Diagrammatic representation of a generic Czerny—Tumer spectrograph, 2001 by Taylor & Francis Group, LLC focal length for a given spot size, but then the system size and cost increases and the system must then be scanned to provide the same wavelength coverage. ‘The principal advantages of the Czerny—Turner are its relative simplicity and wave- length flexibility. Because it is an all-reflective design, it is naturally achromatic. The Czerny—Turner is easy to scan and can accommodate multiple gratings. This makes the Czerny—Turner an attractive spectrograph for research systems that require flexibility to operate with many different lasers (e.g., for resonance Raman spectroscopy [99]). Disadvantages of the Czerny—Turner design are in the areas of etendue, imaging performance, durability (as described in Section ILC. 1.a), and thermal/mechanical stability. ‘Vignetting at the edges of the spectrum is relatively large due to the large distance between the output collimator and the reflection grating. Thermal instability of a Czerny-Turner spectrograph is about a factor of 5 higher than that of a similarly configured axial trans- missive spectrograph based on volume-phase holographic gratings [126]. ‘The use of a reflection grating in the Czerny—Turner design forces the user to make a choice between usable simultaneous spectral range and resolution because this design is unable to produce a two-dimensional (2D) spectral dimension at the focal plane. In order to simulate this, it has been proposed that the input aperture be shared and two gratings be used instead of one (Fig. 8). In this arrangement, the two gratings are set at different input angles with respect to the incoming radiation so that the images are separated in a direction perpendicular to the wavelength dispersion. The spectrometer designer chooses to reduce the optical throughput of the spectrometer by 50% by dividing the system aperture into two. Two physically separated diffracted images are observed at the focal plane. The obvious drawback to the aperture-sharing approach for Raman spectroscopy is that because Raman scattering is a very weak process, it is less than desirable to reduce the optical thoughput of the system. b. Concave Holographic. A concave holographic-grating spectrograph is a single- element spectrograph [127]. A reflective grating is formed on a curved focusing surface to yield an off-axis spectrograph, as shown in Fig. 9. The grating is typically a holographic Grating 1 Diffracted radiation i Incoming collimated iZ L radiation [-——~ Range 1 ss DSSS Range 2 Grating 2 T CCD detector (Focusing concave mirror not shown) Figure 8 Schematic of a Czeny—Turner spectrograph utilizing “aperture sharing” and two sep- arate reflection gratings. By angling grating | with respect to the incoming radiation and grating 2, two different spectral ranges can be observed at the CCD detector. Note that the focusing mirrors are not shown. Copyright © 2001 by Taylor & Franeis Group, LLC Entrance slit Exit slit Exit slit Figure 9 Diagrammatic representations of two different concave holographic spectrographs. surface-relief pattern (or replication thereof) with a reflective overcoat. This is the simplest possible spectrograph design and, therefore, is very compact and inexpensive. Because the grating and imaging elements are one and the same, this design has basically no vignetting problems and can be designed to operate at low fi#. It is, however, more susceptible to image quality problems over any extended wavelength range. Durability and stability are similar to other reflective designs. Many commercially available spectrographs use this configuration, including examples from JY Horiba [123], Oriel [128], and Ocean Optics [129]. c. Reflective/Refractive Hybrid. A reflectivelrefractive variant on the Czerny- Turner has been developed. In this configuration, the mirrors of the Czerny—Turner are replaced with lenses and operated on axis. The use of lenses provides a significant amount of flexibility to improve the image quality away from the optical axis. Because the grating is reflective, the ability to scan is available. The design suffers from the vignetting dis- advantage of reflective designs, exacerbated by the bulk of multielement imaging lenses, as well as the durability and stray-light issues associated with reflective gratings. The vignetting issue makes it difficult to employ fast (low j/#) lenses; therefore, available versions of reflective/refractive hybrids have long focal lengths and high (slow) fi#’s com- parable to Czerny—Turner spectrographs. This approach is somewhat better than an all- reflective Czerny—Turner design for stability due to the reduced number of reflective elements, but still more sensitive than an all-transmissive design. 4. Echelle. An echelle grating and the echelle spectrometer have some potential attraction for multidimensional analytical spectroscopy [130]. The echelle grating is a very low-frequency (less than 100 lines/mm) reflective grating. The major implementation for Copyright © 2001 by Taylor & Francis Group, LLC echelle spectrometers is in atomic emission spectroscopy [131,132]; however, there have been some notable Raman applications of echelle-based instruments [50,133,134]. In Raman applications, the input collimated light is directed to a cross-dispersion element whose function is to provide dispersion in a plane perpendicular to that of the echelle grating. It should be noted that both a grating [133] or prism(s) [134] can be used for this purpose (Figs. 10 and 11). The diffracted light is then passed to an echelle grating at a very low “grazing” angle. The echelle grating is constructed to diffract the light into multiple orders (nine or more) with reasonably high efficiency. Each of these orders rep- resents a different slice of the total wavelength range. ‘The echelle spectrometer then makes use of a two-dimensional detector, as each order occupies a different row of elements, in effect synthesizing a detector much wider than it actually is. Thus, an echelle can cover a wide wavelength range at high resolution (1-2 cm”! design dependent). Pelletier {133} reported the use of nine orders producing 450 cm”! per row, whereas in the Carrabba et al. design [134], more orders were used with approximately 300 cm“ per row. Commercial designs for application in process Raman have been reflective/refractive hybrids employing lenses as collimators. An example is a system offered by InPhotonics [135], which has been successfully used for forensic measurements both in the laboratory and for remote field measurements. The advantage of this type of spectrometer when compared to standard concave ho- lographic or Czerny—Tumer designs are as follows: 1. No moving parts 2. Wide spectral coverage 3. A large reduction in time when collecting a complete spectrum Compact, rugged instrument footprint Improved imaging performance due to refractive optics ae Efficient coupling with fiber-optic devices due to fast refractive optic The first drawback of the echelle design is the difficult problem of properly stitching together the multiple “subspectra” into a properly calibrated continuous spectrum (Fig. igure 10 Diagram of the echelle spectrograph developed by Pelletier (133]: SI, entrance slit; LI, collimating lens; MI, mirror; CD, cross-disperser (grating); EG, echelle grating: L2, focusing lens; FP, spectrograph focal plane. (Adapted with permission from Ref. 133.) Copyright © 2001 by Taylor & Franeis Group, LLC STRAY LIGHT -RTURE, FIBER OPTIC hs me \ 180mm FOCAL LENGTH PRISM, Ny a8) i \ Me — cep DETECTOR, FOCUSING LENS ECHELLE 180mm FOCAL LENGTH GRATING 2.8) 526 grimm STRAY LIGHT ‘APERTURE FIBER OPTIC COLLIMATING LENS 150mm FOCAL LENGTH 3.0) — cop DETECTOR ECHELLE FOCUSING LENS GRATING 150mm FOCAL 526 grimm LENGTH (22.5) (>) Figure 11. Schematic representations of the (a) visible and (b) NIR versions of a commercial echelle spectrograph manufactured for Raman spectroscopy. (Adapted with permission from InPhotonic, Inc.) Copyright ©2001 by Taylor & Frais Geaup, LLC 2.4: > 1.8 a 3 5 1.2 3 s 60 -3300 -2400 -1500 -600 300 Wavenumber from 514.5 nm Figure 12. Calibration spectrum used for the CCD echelle spectrograph. The lower curves rep- resent the white-light response for the different echelle orders used. The lines represent atomic emission lines from a neon lamp used for wavelength calibration. (Adapted with permission from Ref. 133.) 12). This places strict requirements on the stability of the system, as each order must be calibrated and joined. The procedure for calibration is similar to that employed for other spectrographs, but the added difficulty is that there are many different orders (spectral regions) to be joined, An additional problem is vignetting, which can be especially severe for an echelle-based spectrograph, because the grazing angles and cross-dispersion optics keep the collimators a large distance from the grating. The third drawback is that, with current designs, it is not possible to use either a nonimaging probe or multiple-imaging fiber-optic probes (multichannel capability) with this design. The reason for this is that multiple orders from multiple fiber-optic inputs will overlap at the CCD due to the high number of orders used in this configuration e. Axial Transmissive. The axial transmissive spectrograph is one of the newest design concepts commercially available. Originally developed at Kaiser Optical Systems, Inc., [91], this design makes use of lenses for collimators and uses volume-phase holo- graphic transmission gratings, as shown in Fig. 13. This spectrograph design evolved specifically to address the unique requirements of fiber-coupled process Raman spectros- copy [118,136]. The transmission geometry allows the collimators to be placed very close to the grating, enabling the use of commercially available low ii lenses with low vi- gnetting. Units are available as fast as //1.4, which allows these spectrographs to capture all of the light from standard telecommunication fibers without magnification/resolution trade-offs. This translates into a very compact spectrograph package. Typically, off-the- shelf multielement lenses are used, which provide adequate chromatic correction for the Raman spectrum and have excellent imaging quality over a large image plane [109]. Usually, there is less than 5 jum of blurring over a 1-in, detector format. As noted previously, transmissive optics posses a natural stability advantage over their reflective counterparts. To a first approximation, one would expect a transmissive spectro- Copyright © 2001 by Taylor & Franeis Group, LLC Silicon CCD detector chip a Focusing lens (f/1.4) Entrance sit Collimator lens (111.8) rT ‘Holographic VPT grating TD Figure 13 Schematic representation of an 11.8 axial transmissive spectrograph. (Reproduced with permission from Kaiser Optical Systems, Inc.) graph to have at least a 4 to 1 advantage in calibration and alignment stability. In practice, the advantage has been shown to be 5 to 1 or more [126]. Transmission gratings can be manufactured with a high diffraction efficiency (>80% efficiency for unpolarized light). Because the grating is transmissive and diffracts only a limited range of wavelengths, a unique opportunity for multiplexing exists. One can in- clude a second grating designed to diffract a different wavelength range into the same sealed package. The two gratings are tilted with respect to each other, thus directing the two wavelength ranges to two different vertical portions of the detector [115,116]. In so doing, a detector twice as wide is synthesized, similar to the multiple diffraction orders of an echelle, Because the efficiency profiles of the transmission gratings are much smoother and more stable than in an echelle, it is much less difficult to accurately recon- struct the continuous spectrum. ‘An alternative transmission grating approach has been proposed by de Castro Carranza et al. [137]. In this approach, the two gratings are recorded into a single hologram film rather than in multiple hologram layers. This approach, however, is not recommended for Raman spectroscopy, as the effective modulation of each grating is reduced with respect to individual holograms. Thus, in a method analogous way to aperture sharing in Czerny— Turner spectrometers, the net throughput of the system is reduced. The advantage of the multitransmission design when compared to standard concave holographic or Czerny—Turner designs are the following: 1. No moving parts 2. Complete Raman spectral coverage 3. Allows a large reduction in time when collecting a complete spectrum Compact, rugged instrument footprint Copyright © 2001 by Taylor & Francis Group, LLC 5. Improved imaging performance due to refractive optics 6. Low vignetting due to the close proximity of the grating to the output lens 7. Efficient coupling with fiber-optic devices due to fast refractive optics 8 Can be user optimized for either a specific or several excitation wavelengths by swapping the grating and refocusing the len: In the process environment, the user is generally interested in a single application- specific excitation laser wavelength. Thus, the axial transmissive spectrograph design is particularly well suited to process Raman applications. The principle disadvantage of the axial transmissive design is in wavelength flexibility. ‘The low /f# lenses used can be chromatically corrected only for a limited wavelength range. This correction is enough to cover the Raman spectra of a given laser at high resolution, but requires refocusing of the lens for different excitation laser wavelengths if optimum resolution is to be retained. Similarly the gratings are inherently wavelength range limited and are not amenable to scanning. To change the wavelength and range for a different excitation wavelength, the gratings must be changed, as well as refocusing the spectrograph lenses (a fairly trivial process). This disadvantage can be offset, as the com- bination of refocusing the spectrometer for a specific excitation wavelength range and using a grating optimized for a given excitation wavelength yields a high-throughput spectrometer: fi Fixed Fitter/Reflection Grating. A new variant of the reflective/refractive spec- trometer has recently been developed which allows for two or more dimensions of si- multaneous spectral coverage [138-140]. A schematic of the spectrometer is shown in Fig. 14. In this device, the incoming light is first collimated and then passed to a series of cutoff or edge filters. The function of these filters is to separate the incoming light into X broad wavelength ranges by reflecting a predefined wavelength range and then trans- G1 uu RL MI ccD Incoming collimated radi: Mion Figure 14 Diagrammatic representation of the fixed filter/grating spectrograph recently reported by Smith [1995]: Fl, filter range 1; F2, filter range 2; MI, mirror 1; Gl, reffection grating; L1, CCD focusing lens; R1, range I relates to radiation reflected by F1; R2, range 2 relates to F2; R3, range 3 relates to MI Fl Copyright © 2001 by Taylor & Franeis Group, LLC mitting the remainder to the next filter (X is the number of cutoff or edge filters). The reflected light is diffracted by a grating and imaged as X physically separate strips of the CCD detector. Smith et al. [138] mostly describe a system based on HRDGs, but an embodiment using a holographic transmission grating is also shown. This spectrometer is claimed to have many of the advantages of the echelle and axial transmissive designs. The authors do note some disadvantages, including the following: 1. The need for sharp cutoff or edge filters. The possibility of lens aberrations limiting the spectrometer performance. 3. The potential for vignetting due to the fact the final focusing lens must be located away from the HRDG. 4, Changing excitation wavelengths is complex. In the approach reported by Smith et al. [138], the grating angle must be rotated, multiple filters must be changed, and the lens optics must be recollimated, For process measurement where a single excitation wavelength is used, this spectrometer may be considered. g. Tunable Filter, An alternative approach to spectrographs for measuring Raman spectra is to use one of the tunable-filter technologies available; dielectric interference [141], acousto-optical (AOTF) [142], and liquid-crystal (LCTF) [143]. These devices are very compact and rugged, but are predominantly used as tunable single-pass-band devices for high-fidelity Raman imaging (and, as such, are discussed in those terms in depth in Chapter 5). It is unlikely that these devices will be used for process Raman spectroscopy, as only a single wavelength can be collected at any one time. These devices are not effective when studying temporally changing signal such as in turbulent media (in a man- ner similar to FT spectrometers and scanning spectrographs). These devices, in conjunction with a single-channel detector, could be used to produce a simple Raman spectrometer. The resulting instrument could be used to study single discrete vibrational bands if these were the only required feature in the process Raman measurement, The cost, however, of the AOTF and LCTF devices is likely to make this, approach prohibitive. D. Detectors Because Raman scattering is such a weak effect, detector performance is critical. Ideally, the detector should operate in a shot-noise-limited regime. Shot noise is a fundamental, theoretical limitation on the noise performance of any photon detector. It stems from the statistical uncertainties in the arrival times of detected photons. In the ideal detector, all other noise sources are at levels well below the shot noise. The shot noise is the square root of the number of detected photons. For applications with lots of available light, such as absorption spectroscopy, even a detector with poor inherent noise performance may be operating in a shot-noise-limited regime. The Raman effect leads to the production of very few photons (1 Raman photon for a strong Raman scattering material for every 10° incident photons). In order to collect enough photons to yield a good signal-to-shot-noise ratio, long integration times may be necessary. In process-control applications, integration times may range from less than a second up to several minutes. In order for these weak signals to remain shot-noise limited, the other noise contrib- utors of the detector must be sufficiently low. Most significant is dark current or dark 2001 by Taylor & Francis Group, LLC noise, which is a process where electrons that are thermally generated in the detector material accumulate in the detector storage medium even with no light falling on the detector. When the detector is read out, these thermally generated electrons are indistin- guishable from those generated by photons of light. Other sources of noise include readout noise, quantization noise, and background or stray-light noise. Read noise is due to the readout electronics of the detector and is basically inaccuracy in measuring the number of electrons. Quantization noise relates to the resolution of the analog-to-digital conversion of the readout signal. Background or stray-light noise is due to light hitting the detector that is not really part of the signal being measured and as such is a property of the overall stray-light characteristics of the optical system, as well as the effectiveness of the method of laser-line rejection. Single-clement detectors can be used for Raman spectroscopy in wavelength-scanning spectrographs. A multielement detector, however, is more desirable because it allows si- multaneously collection and integration of light over a large spectral range. A two-dimen- sional array detector is better still, as this allows instantaneous collection from multiple sources (fibers) and/or multiple wavelength ranges. Ideally, the detector should have a large number of closely spaced elements, wide dynamic range, wide wavelength range, and high quantum efficiency. Quantum efficiency (QE) is the percentage of incoming photons that are converted into detectable photoelectrons. In this case, more is simply better, all else being equal. In general, more elements are better, but the maximum usable width of the detector is limited by the available image plane of the spectrograph imaging optics (this is approximately | in. for current systems). This leads to the desire for smaller elements, but only a few spectrograph designs have the image precision to make effective use of them, as discussed in Section II.C.1.e. It should also be noted that there is at some point a trade-off between the size of the detector element and its charge storage capacity. This will be discussed in more detail subsequently. ‘The dynamic range of a detector is also important. The dynamic range is simply the difference between the maximum and minimum detectable signal levels. At first glance, one would think a detector with eight orders of magnitude of dynamic range would be appropriate for a Raman system, as that is roughly the difference between the laser inten- sity and the Raman signal. However, as covered in Section I1.C.1.4, the stray-light ratio of most practical spectrographs is closer to four or five orders of magnitude, limiting the dynamic range of the system. Although more is preferable to less, anything much above the spectrograph stray-light performance is essentially wasted at the detector level. In fact, the primary function of the laser-line rejection device is to compress the required dynamic range of the spectrograph and detector combination to achievable values. ‘The final consideration when selecting a detector is the wavelength range of interest. This depends both on the excitation laser(s) used and on the location of the bands of interest. A truly universal Raman detector would cover the entire range of readily available lasers, fibers, and imaging optics, plus the full range of Raman shift on the long-wave- length side. In the process industry, this range can be restricted to that where diode or diode-pumped lasers and standard fiber technology can be used. This wavelength range is thus from about 400 to around 2000 nm (or in terms of absolute frequency, from 25,000 to 5000 em”). 1. Photomultiplier Tube The photomuttiplier tube (PMT) was the first modern electronic detector available for Raman spectroscopy [144]. The coupling of a PMT and a scanning monochromator al- Copyright © 2001 by Taylor & Franeis Group, LLC lowed the whole Raman spectral range of interest to be measured, albeit one resolution element at a time. The advent of multichannel detectors for Raman spectroscopy, initially the vidicon, subsequently the diode array, and most recently the CCD array, has signifi- cantly reduced the popularity of the PMT by allowing polychromators to be produced. The reasons for the relegation of the PMT from its historical position as detector of choice to its current position are as follows: 1, The time required to measure the spectrum with a scanning monochromator 2. The wavelength response of PMTs, which is significantly reduced for both red (632.8 nm) and NIR (785 nm) excitation when compared with green excitation (514.5 or 532 nm) 3. The susceptibility to damage of the active photocathode (when exposed to intense radiation) 4, The high-voltage requirements of the device and the photon-counting electronics required 5. The need for water cooling to reduce dark noise 6. The fragile nature of the device itself For these reasons, PMTs are now rarely used in the laboratory setting and, in general, are unsuitable for process Raman spectroscopy. The one case where PMTs may be useful for process analysis is when a single discrete band of interest needs to be measured. In this case, a filter-based process Raman instrument could be produced with a PMT detector to construct a relatively low-cost instrument. There have been some reports in the literature where PMTs have been used for process Raman spectroscopy. These include studies of natural gases by Florrison and Burrie [145] and anesthetic respiratory gas by employees of Albion Instruments [146,147]. For more information on these applications, the reader is referred to Chapter 23. The above-mentioned reports are more than 10 years old, written, at a time when the PMT was still considered a state-of-the-art detector. 2. Charge-Coupled-Device Array Although other detector technologies exist, the current detector of choice for virtually all types of dispersive Raman spectroscopy is the silicon CCD (charge-coupled device) array. The CCD array meets more of the desired detector characteristics for Raman spectroscopy than any other currently available detector technology. These characteristics include the following: 1. The ability to measure many wavelengths at once (up to the full spectrum de- pending on spectrograph configuration) 2. Large wavelength range 400-1000 nm 3. Large dynamic range 4, High quantum efficiency 5. The ability to store and move charge 6. Low read noise 7. Low dark noise when cooled by either water exchange, thermoelectric, or cryo- genic means 8. Durability when exposed to intense radiation 9. Robust packaging and long detector lifetime Copyright © 2001 by Taylor & Francis Group, LLC Charge-coupled devices are commonly available in a wide range of formats. Most useful for spectroscopy are CCDs whose shapes are rectangular (1 in. wide by 0.25 in, tall is a typical form factor) because this best matches the focal plan of the spectrograph. a. Operation, There are a number of excellent references [131,148—151] that dis- cuss the fundamentals of CCDs; thus, only a cursory discussion will be given here. CCDs ‘on devices constructed with lithographic fabrication techniques similar to those used in computer-chip manufacture. The silicon is doped and a series of electrodes are deposited on the surface (Fig. 15). Because silicon has a high reflectivity, the surface of the device is often treated with an antireflection coating. Voltages can be applied to these electrodes by specially designed electronics which produce a two-dimensional array of potential wells just under the surface of the silicon; these wells are referred to as the pixels of the CCD. The wells will attract and hold electrons generated in the silicon by either heat or light. Exposing the CCD to light will cause electrons to build up in each charge well at a rate proportional to the intensity falling on that area. Each pixel well of a CCD typically used in a Raman process system will hold around 200,000 electrons. This two- dimensional detector is composed of what are termed rows and columns of pixels. Output Read-out Register amplifier @ 1024 x 256 active area CCD array o1 02 — 1 03 Single Pixel (b) Channel Stop Figure 15 (a) Schematic of a typical spectroscopic CCD chip; 1024 x 256 front-illuminated MPP chip; (b) Expanded view of the electronic architecture of a typical spectroscopic three-phase CCD detector. Copyright © 2001 by Taylor & Franeis Group, LLC The CCD is read out by sequentially changing the voltage level of the electrodes for cach row of wells on the detector (Fig. 16). This causes the electrons from one row of wells to shift into an adjacent row, whose electrons are, in turn, shifting to the next row. In this way, the wells form a sort of bucket brigade, shifting their electrons, row by row across the surface of the silicon, At one edge of the CCD is a final row of wells called the shift register. Each successive shift of the main detector dumps a full row of electrons into the shift-register wells. At one end of the shift register is an amplifier. The register shifts its electrons in the same fashion serially toward the amplifier, which produces a voltage proportional to the number of electrons in each well dumped into it, Finally, the amplifier is connected to a digitizer that produces a digital number proportional to the voltage produced by the amplifier, in effect counting the electrons in each well. In general, scaling (termed gain) is applied by the digitizer, with each count representing several electrons, These numbers are then transferred to the computer to be presented as a spectrum, b. Noise Characterization and Optimization. The noise performance of CCD de- tectors is one of the major features that makes these detectors attractive for Raman spec- troscopy. Many other established detection technologies are currently unsuitable for Raman spectroscopy due to noise performance (e.g., diode arrays and charge-injection devices [152}). For scientific-grade CCDs, the read noise is typically less than 10 electrons, which, for most systems, corresponds to around one or two counts per read event. The number of read events can be varied by how many rows’ worth of electrons are dumped into the shift register before they are shifted to the amplifier. For most Raman systems in process environments, column pixels are combined or “binned” [153,154] such that each wave- length channel, which may subtend a number of rows, requires only one read event. Thus, 01=0V =4V f 03 = 0V—>+V. —— Charge Movement Figure 16 Diagrammatic representation of charge transfer in a CCD detector showing the use of applied voltage to move charge across a pixel. Copyright © 2001 by Taylor & Francis Group, LLC for signal levels that generate at least a few hundred or more counts (which will represent only a few thousand photons), the read noise is not significant. Charge-coupled devices also can have a very low thermal or “dark” count/noise. A nonscientific consumer-grade CCD, not intended for spectroscopy, can generate more than 30,000 electrons/pixel/see at room temperature. It is not uncommon for collection time in the process Raman environment to be on the order of 5 min or more. In this time, the wells of such a CCD would be completely filled just by the dark current. The CCD dark count is reduced through the application of two techniques: multipin phasing (MPP) and cooling. MPP is a design architecture of the CCD chip that reduces the natural dark count rate [154]. MPP is a combination of special processing of the surface insulation layers during manufacture, a special electrode arrangement, and a particular voltage sequence during operation. The net effect is to prevent many of the thermally generated electrons from getting drawn into the potential wells. The use of MPP architecture can reduce the base dark count rate of a CCD chip by a factor of between 50 and 200. An alternative nomenclature for MPP is AIMO (advanced inverted-mode operation). Cooling is fairly straightforward. ‘The dark current of a CCD roughly doubles for every 6°C increase in its temperature. Thus, colder CCDs generated lower the dark nois Originally, spectroscopic CCDs were cooled with liquid nitrogen, which reduces the dark count to essentially undetectable levels even if the chip does not employ a MPP architec ture, Liquid nitrogen, while merely inconvenient for the research laboratory, is simply not acceptable in systems intended for the process environment. There are examples, primarily in the military, of CCDs being cooled with gas-cycle refrigerators such as sterling engines. This technology can approach cryogenic temperatures well below —100°C, but as of this writing, it is not ready for broad commercialization. It is the advent of solid-state ther moelectric (Peltier) coolers that have allowed CCDs to be used for process Raman, High- performance Peltier coolers coupled with the proper vacuum-sealed enclosure can reduce the temperature of the CCD to as much as 70-90°C below ambient. These coolers are typically configured to require only electricity and air for heat dissipation. They have no moving parts and are very reliable. Many commercially available cameras employ a liquid coolant to dissipate waste heat as an option. Although this is considerably more efficient than air, a coolant flow can be a maintenance or logistics problem in a process environment and is generally avoided. The combination of cooling and MPP can reduce the dark current to around 0.01-0.5 electrons/pixel/sec. At this level, the dark count for a 5-min exposure will be comparable to the read noise and not significantly affect the signal. ©. Other Performance Considerations. The dynamic range of CCDs is generally expressed in the number of bits produced by its digitizer. State-of-the-art CCDs have 16- bit digitizers, which gives them a dynamic range of 65,000 counts or about 4.5 orders of magnitude. This is a fairly good match to the stray-light performance of current spectro- graphs and is more than adequate for most process applications. Some CCD manufacturers advertise 18-bit digitizers, which should, in theory, give a dynamic range of over 200,000 counts. In most cases, the underlying precision of the amplifier and other electronics cannot support this range, so, in reality, the extra two bits are mostly noise. The scaling, or gain, of the CCD is the number of electrons required to produce a digitized count. The well capacity of a modern MPP-modified CCD will be around 200,000 electrons. The gain of the CCD is usually set such that the well capacity for each wave- length channel is matched to the range of the digitizer. Typically, several pixels are com- Copyright © 2001 by Taylor & Franeis Group, LLC bined together for each wavelength channel. Gains are normally between 4 and 10 elec- trons per count, which requires between 2 to 4 pixels be combined to have enough electrons to use all 65,000 counts. The operational wavelength range of the CCD is the span of wavelengths over which the quantum efficiency of the detector is usable. This range is from about 400 nm to about 1000 nm as limited by the silicon detection medium and the detector coating. Within this range, the maximum QE for a standard CCD is around 40%. This limited range is the one major weakness of silicon CCDs, placing a hard upper limit on usable excitation wave- lengths. Desirable lasers such as diode-pumped Nd: YAG operating at 1064 nm cannot be used with CCDs (there is the possibility of detecting anti-Stokes Raman but this is gen- erally too weak for process work) [155,156]. An alternative type of CCD is the back-thinned or back-illuminated CCD, in which the chip is flipped upside down and the silicon-active medium is reduced in thickness such that light enters the silicon from the back side, opposite the light-obscuring electrodes. ‘These CCDs have QEs approaching 80% (Fig. 17). The disadvantage of using these types of detectors for 780-785-nm excited NIR Raman spectroscopy (780-1050-nm spectral range) is detector etaloning [94]. In order to increase CCD sensitivity in this range, a different type of CCD using deep depletion silicon has been produced. In these devices, the doping of the silicon is changed such that the resistivity of the silicon is higher. This allows the absorption for incoming photons to occur deeper in the device [154]. The enhanced NIR sensitivity of front-illuminated deep depletion CCDs is shown in Fig. 17. Most recently, a new detector which combines both deep depletion and back-thinning has been produced [157]. In these devices, a proprietary surface treatment is incorporated into production, which effectively eliminates etaloning in the NIR. The NIR performance is compared to alternative CCD devices in Fig. 17. It can be seen from these curves that the back-thinned, deep depletion CCD is the preferred device for NIR-excited Raman spectroscopy. The only drawback to the deep depletion devices is that they tend to be 100) g 8 PCC o A | | © | rSeS 242 Bas 2 & Quantum Efficiency (%) a & | xg 8 8 3 400 «500 «G00 «700 «G00 «800 «1000 ‘Wavelength (nr) Figure 17 Quantum-efficiency curves for a several CCD detectors. (a) Front-illuminated MPP chip; (b) back-illuminated MPP chip; (c) NIR-enhanced front-illumined, deep depletion design— non-MPP chip; (d) NIR-optimized back- illuminated, deep depletion design—non-MPP chip. Copyright © 2001 by Taylor & Francis Group, LLC

You might also like