Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the...

27
30 Raman Spectroscopy: From Benchtop to Bedside 30.1 Introduction 30.2 Principles of Raman Spectroscopy 30.3 Instrumentation Considerations Basic Instrumentation 30.4 Effect of Sample Conditions Tissue Model Selection • In Vivo Considerations 30.5 Processing Calibration • Fluorescence Elimination • Other Preprocessing Methods 30.6 Analysis Automated Diagnosis • Component Extraction 30.7 Clinical Application of Raman Spectroscopy Breast • Cervix • Skin • Other Applications 30.8 Future Perspectives References 30.1 Introduction Light can react with tissue in different ways and its response can yield information about the physiology and pathology of tissue. The light used to probe tissue does so in a nonintrusive manner; in contrast to therapeutic applications, very low levels of light are typically used. The use of fiber optics allows light to probe tissue in a minimally invasive manner. Because tissue response is virtually instantaneous, the results are obtained in real time, and data processing techniques and computers allow for automated detection of disease. These factors have resulted in a variety of applications that employ light and optical modalities for tissue diagnosis. When a photon is incident on a molecule, it may be transmitted, absorbed, or scattered. The various techniques that are based on these different light–tissue interactions include: Absorption spectroscopy Reflectance spectroscopy Fluorescence spectroscopy Raman spectroscopy Each of these four techniques has been studied for the purpose of tissue diagnosis with varying degrees of success. Of these techniques, fluorescence spectroscopy is perhaps the most researched one for tissue diagnosis. Anita Mahadevan-Jansen Vanderbilt University Nashville, Tennessee ©2003 CRC Press LLC

Transcript of Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the...

Page 1: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

30Raman Spectroscopy:

From Benchtopto Bedside

30.1 Introduction30.2 Principles of Raman Spectroscopy30.3 Instrumentation Considerations

Basic Instrumentation

30.4 Effect of Sample ConditionsTissue Model Selection • In Vivo Considerations

30.5 ProcessingCalibration • Fluorescence Elimination • Other Preprocessing Methods

30.6 AnalysisAutomated Diagnosis • Component Extraction

30.7 Clinical Application of Raman SpectroscopyBreast • Cervix • Skin • Other Applications

30.8 Future PerspectivesReferences

30.1 Introduction

Light can react with tissue in different ways and its response can yield information about the physiologyand pathology of tissue. The light used to probe tissue does so in a nonintrusive manner; in contrast totherapeutic applications, very low levels of light are typically used. The use of fiber optics allows light toprobe tissue in a minimally invasive manner. Because tissue response is virtually instantaneous, the resultsare obtained in real time, and data processing techniques and computers allow for automated detectionof disease. These factors have resulted in a variety of applications that employ light and optical modalitiesfor tissue diagnosis.

When a photon is incident on a molecule, it may be transmitted, absorbed, or scattered. The varioustechniques that are based on these different light–tissue interactions include:

• Absorption spectroscopy

• Reflectance spectroscopy

• Fluorescence spectroscopy

• Raman spectroscopy

Each of these four techniques has been studied for the purpose of tissue diagnosis with varying degreesof success. Of these techniques, fluorescence spectroscopy is perhaps the most researched one for tissuediagnosis.

Anita Mahadevan-JansenVanderbilt UniversityNashville, Tennessee

©2003 CRC Press LLC

Page 2: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

Considerable evidence indicates that fluorescence spectroscopy of exogenous (see, for example, Ref-erence 1) and endogenous chromophores can be used to identify neoplastic transformations in the breast,2

lung,3 bronchus,4 oral cavity,5 cervix6 and gastrointestinal (GI) tract.7,8 Despite the success of this tech-nique, results indicate that fluorescence spectra of precancerous tissues of the cervix and colon and benignabnormalities such as inflammation and metaplasia are similar in many patients.6–8 This suggests thatthe use of fluorescence diagnosis in a screening setting, where the incidence of precancer is expected tobe low, may result in an unacceptably high false-positive rate. Many groups consider vibrational spec-troscopy to be the solution to these problems because its molecular specificity enhances the specificityof spectroscopic diagnostics. The focus of this chapter is the application of Raman spectroscopy inbiomedicine, especially as it pertains to disease detection.

Raman spectroscopy has been used for many years to probe the biochemistry of various biologicalmolecules.9,10 In recent years, there has been interest in using this technique in diagnostics.11–14 Most ofthe diagnostic applications are still in the early stages of development. They range from the detection of(pre)malignant lesions in various organ sites to quantitative detection of blood analytes such as glucose,from measurement of skin hydration to the study of aging. The focus of all of these applications is onquantitative, in vivo, nonintrusive, automated detection in a real-time manner.

This chapter describes the concept, instrumentation, and application of Raman spectroscopy as itapplies to biomedicine. We begin with a description of the basic concepts underlying Raman spectroscopyand Raman signatures of materials. We then review the past, present, and future directions of instru-mentation that facilitates the use of this technique for detection of disease. Finally, we examine sampleapplications of Raman spectroscopy for disease diagnosis, with reference to work at the molecular andmicroscopic levels. Although many of the reported articles are based on in vitro work, this chapter focuseson the potential of the application of Raman spectroscopy to clinical diagnosis.

30.2 Principles of Raman Spectroscopy

Raman spectroscopy is based on the Raman effect that results from energy exchange between incidentphotons and the scattering molecules. When the energy of the incident photon is unaltered after collisionwith a molecule, the scattered photon has the same frequency as the incident photon. This is Rayleighor elastic scattering.15,16 During the collision, when energy is transferred from the molecule to the photonor vice versa, the scattered photon has less or more energy than that of the incident photon. This isinelastic or Raman scattering (Figure 30.1).

Quantum mechanically, Raman scattering is a result of an energy transition of the scattering moleculeto a virtual excited state and its return to a higher (or lower) vibrational state with the emission of analtered incident photon (Figure 30.1). Because energy can be transferred from the photon to the moleculeor from the molecule to the photon, the scattered photon can have less or more energy compared to theincident photon. When the scattered photon has less energy than the incident photons, the process isreferred to as Raman Stokes scattering. When the scattered photon has more energy than the incidentphotons, the process is referred to as Raman anti-Stokes scattering.

A Raman spectrum is a plot of scattered intensity as a function of the frequency shift (which isproportional to the energy difference) between the incident and scattered photons (Figure 30.2). Thefrequency shift is characteristic of the molecule with which the photon collided, and the resultingspectrum is characterized by a series of Raman bands that correspond to the various vibrational modesof that molecule. The locations of the Raman bands are typically presented in terms of relative wave-numbers (or Raman shift) defined as shifts in wavenumbers (inverse of wavelength in cm–1) from theincident frequency, which is set to zero. This is because the location of Raman bands for a given moleculein relative wavenumbers remains constant, regardless of the incident frequency (within the physicalconstraints of the light–tissue interaction).

Raman signals are usually weak and require powerful sources and sensitive detectors. Typically, Ramanpeaks are spectrally narrow (a few wavenumbers) and, in many cases, can be associated with the vibrationof a particular chemical bond (or normal mode dominated by the vibration of a single functional group)

©2003 CRC Press LLC

Page 3: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

in a molecule. Consider a complex molecule such as glucose (Figure 30.2). Each of the stretching andbending modes of vibration of this molecule has a characteristic frequency. When a photon is incidenton this molecule, each mode of vibration results in a characteristic shift in frequency that is indicatedin the Raman spectrum. Because each type of bond has characteristic modes of vibration, each type ofmolecule has its own spectral “fingerprint.” Thus, in a complex mixture of molecules such as in tissue,the presence of the unique bands of glucose can be traced; this results in the quantitative evaluation ofthe sample’s chemical composition, which can then be related to the tissue pathology for diagnosis.

Raman spectroscopy probes characteristics of materials other than fluorescence.15 The energy tran-sitions of molecules are solely between the vibrational levels. When a photon is incident on a molecule,it may be transmitted, absorbed, or scattered. Fluorescence results from the emission of absorbed energy;Raman scattering results from perturbations of the molecule that induce vibrational or rotational

FIGURE 30.1 Schematic illustration of classical and quantum mechanical explanation of scattering.

FIGURE 30.2 Raman spectrum of glucose in powdered form obtained with dispersive Raman spectroscopy at 785-nm excitation.

2ν 4ν

S0(0)

S0(1)S0(0)

S0(1)S0(0)

S0(1)

3ν 3ν 3ν 2ν 3ν 4ν

Rayleigh orElastic Scattering

Raman StokesScattering

Raman Anti-Stokes Scattering

0.00E + 00

5.00E + 03

1.00E + 04

1.50E + 04

2.00E + 04

2.50E + 04

500 700 900 1100 1300 1500 1700

Raman Shift (cm−1)

Inte

nsi

ty (

a.u

.)

Glucose

HOCH2

O H

H

H

HOH

OH

HOOH

©2003 CRC Press LLC

Page 4: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

transitions. Only a limited number of biological molecules such as flavins, porphyrins, and structuralproteins (collagen and elastin) contribute to tissue fluorescence, most with overlapping, broadbandemission. In contrast, most biological molecules are Raman active with fingerprint spectral character-istics. Indications are that vibrational spectroscopy may overcome some of the limitations of fluorescencediagnosis of precancers and cancers.

Several different modalities of Raman scattering have been used to analyze the structure of variousbiological molecules.17,18 These techniques include near-infrared (NIR) dispersive Raman, Fourier trans-form Raman (FT-Raman), surface-enhanced Raman (SERS), and ultraviolet resonance Raman (UVRR)spectroscopy.15,16 In NIR dispersive Raman spectroscopy, NIR radiation typically in the range of 780 to1100 nm is used for excitation. The advantage of this technique is that minimal fluorescence is produced,thus making detection of the weak Raman signal easier. In FT-Raman spectroscopy, the Fourier transformof the signal is detected and then inversely transformed to give the actual Raman signature. This techniqueyields improved signal-to-noise ratio (SNR) of hard-to-detect events, but requires high collection times.13

Recent developments in spectroscopic instrumentation have resulted in reduced application of FT-Ramantechniques in tissue diagnosis. Thus, while some early results from FT-Raman studies are referenced, thefocus of this chapter is on dispersive Raman spectroscopy.

SERS is used to investigate the vibrational properties of single adsorbed molecules.15,16 It was discoveredthat the rather weak Raman effect can be greatly strengthened (by a factor of up to 14 orders of magnitude)if single molecules are attached to nanometer-sized metal structures. Metal surfaces must be of highreflectivity and of a suitable roughness. In this way, many groups have detected single molecules attachedto colloidal silver particles adhered to a glass slide or even in an aqueous solution. The advantages of thismethod are that it is fast, it can supply some structural information about the molecules, and it doesnot bleach the molecules. Single-molecule detection is of great practical interest in chemistry, biology,medicine, and pollution monitoring; examples include DNA sequencing and the tracing of interestingmolecules such as those used in bioterrorism. Again, because the focus of this chapter is on clinical invivo diagnosis, SERS will not be discussed here.

When the excitation frequency approaches or enters the region of electronic absorption of a molecule,the resonance Raman spectrum of that molecule is obtained.16 By choosing appropriate excitation fre-quencies, a sample is selectively excited at the maximum absorption frequency of a characteristic moleculeto detect that feature above all else. Resonance excitation increases the scattering intensity by severalorders of magnitude. Typical absorption frequencies of biological molecules such as proteins and nucleicacids are in the ultraviolet portion of the spectrum, where the Raman and fluorescence signatures maybe spectrally resolved. However, the high intensities may cause photolysis of the sample and destroy itover time. Besides, this signal may be attenuated by simultaneous intensified absorption and fluorescenceemission.16 In addition to these factors, the mutagenicity of UV radiation makes this technique inviablefor clinical in vivo use.19 Therefore, UVRR studies are excluded from this chapter.

Many groups have recognized the potential of Raman spectroscopy in the study and diagnosis ofdisease because of its successful application in biology. Early attempts to measure Raman spectra ofcells and tissues were hindered by two factors: (1) the highly fluorescent nature of these samples and(2) instrument limitations, which necessitated long integration times and high-power densities toachieve spectra with good S/N ratios. Improvements in instrumentation in the last decade, particularlyin the near-infrared region of the spectrum where fluorescence is reduced, have resulted in a dramaticincrease in biomedical applications of Raman spectroscopy. Recent reviews of this field11,20–22 illustratethe diversity of potential applications, ranging from monitoring of cataract formation in vivo23 to theprecise molecular diagnosis of atherosclerotic lesions in coronary arteries.24,25 Reports have begun toappear validating the potential of Raman spectroscopy to diagnose disease in human tissues in vivo.36,27

30.3 Instrumentation Considerations

Despite the wealth of information provided by Raman spectroscopy about the structure of biologicalmolecules, early attempts to measure Raman spectra of tissues were limited by two factors: (1) the highly

©2003 CRC Press LLC

Page 5: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

fluorescent nature of these samples and (2) instrument limitations, which necessitated long integrationtimes and high-power densities to achieve spectra with good S/N ratios. The initial Raman spectra oftissue were measured with visible laser excitation, using primarily the argon laser lines (see, for example,Yu and East28 and Clarke et al.29). With the development of interferometers, FT-Raman spectroscopy wasused to measure tissue Raman spectra, typically using 1064 nm (Nd:YAG) for excitation with germaniumdetectors.10

The development of diode lasers and cooled silicon CCD cameras sensitive in the near-IR has resultedin a wide range of applications beyond the laboratory. Diode lasers can provide excitation in the regionof 750 to 850 nm, which allows the use of silicon detectors (sensitive only to 1100 nm). There are twoadvantages of this technique: (1) fluorescence emission is reduced and (2) spectra with acceptable SNRsratios can be achieved with relatively short integration times on the order of a few seconds. Thus, Ramanspectroscopy is now commonly used in environmental monitoring and manufacturing. Advances in theRaman instrument by the commercial sector that supplies this industry have positively impacted theapplication of Raman spectroscopy in biomedicine, making clinical use viable.

30.3.1 Basic Instrumentation

The basic instrument capable of measuring Raman spectra is similar to any spectroscopic system. Itconsists of a light source, which is typically a laser, light delivery and collection, and a detection system.Figure 30.3 shows a typical clinical Raman system used today.

The weak nature of the Raman phenomenon makes the details of this basic system challenging.These three basic components interact in terms of overall system performance, but as a first approx-imation they can be viewed as individual components contributing to the system. Significant techno-logical advances in each of these areas have resulted in Raman instruments superior to similar systemsfrom the previous decade. We consider each of the three components of this basic system individually;however each is dependent on the intended use of the system. For example, a system capable ofconfocality requires a single-mode laser and appropriate optics. A clinical fiber-based Raman systemrequires a multimode diode laser with stable frequency output and rugged design that facilitatesportability.

FIGURE 30.3 Schematic of a typical near-infrared dispersive Raman spectrometer used for tissue diagnosis today.This system is shown with an Enviva Raman probe.

SAMPLE

Collection Fibers

Spectrograph

LN-CCD Camera

λ = 785 nm300 mW

Computer

Diode Laser

©2003 CRC Press LLC

Page 6: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

30.3.1.1 Light Sources

Traditional Raman systems used the argon ion (Ar+) laser for visible excitation, the ND:YAG laser forFT-Raman applications and the Titanium:Sapphire (Ti:Sapph) laser for NIR excitation due to their highoutput powers, single spatial and longitudinal mode operation, and Gaussian beam profile that allowsfor near-diffraction-limited optical performance.30–33 However, the size and electronic and coolingrequirements of these lasers make them impractical in a portable clinical system. Some current Ramaninstruments, especially those with confocal capabilities, still use the Ti:Sapph laser. However, advancesin diode laser technology have completely changed the fingerprint of a typical Raman system.

Diode lasers utilize electro-optical components (diodes), which emit light as a function of appliedcurrent and operating temperature. The laser diodes are small (<1 mm3) and require highly accuratecontrolling electronics to obtain stable output. Laser diodes require highly stabilized temperature con-trollers to minimize thermoelastic effects on the laser cavity length (and thus output frequency), andhighly stabilized current sources to minimize output power fluctuations. However, even with highlyaccurate controlling electronics, laser diodes are still susceptible to output variations, so users of theselasers must be aware of their limitations. In addition to the frequency and power considerations described,laser diodes are also characterized by their elliptical beam output (due to the rectangular shape of theoutput facet) and astigmatism (due to unequal beam divergence from each dimension of the rectangularfacet). These issues also need to be accounted for when using a diode laser. Most commercial diode lasers,however, are available with a pig-tail option where a fiber is coupled directly to the laser diode, thusminimizing losses due to astigmatism and elliptical nature of the beam.

A diode laser with tunability in its wavelength is available; such a system tends to be more susceptibleto changes in the laser cavity length, resulting in so-called “mode-hops” in which the primary resonantmode of the laser jumps from one frequency to another. Because Raman scattering is very sensitive tothe frequency of the incident photon, a source stable in frequency with high line widths is essential toobtain reliable spectra. Diode lasers that are optimized for highly stable frequency output and powers ofthe order of 300 mW specifically designed for Raman spectroscopy are now commercially available (suchas that offered by Process Instruments Inc., Salt Lake City, UT).

Most published reports of confocal Raman microspectroscopy utilize bulky Ar+ and Ti:Sapph lasersfor excitation sources.34 However, the size of these lasers, and their electronic and cooling requirement,make use in a portable clinical system impractical. Thus, in order to develop a truly portable confocalRaman microscope, an alternative excitation source is needed. By extending the length of the resonantcavity of a laser diode, the distance between the diode’s longitudinal modes can be extended, therebyminimizing the effect of small thermoelastic changes on the output frequency. Such a source is calledexternal cavity diode laser (ECDL). Compared to a bare laser diode, the ECDL nearly eliminates mode-hops, minimizes spectral bandwidth of the output light, allows substantial wave length tunability, andmarkedly decreases frequency dependence on temperature stabilization. A commercial ECDL single-mode system, previously available at a high price, is no longer on the market. An ECDL excitation sourcecan be built in-house using off-the-shelf optics, hardware, and electronics at a price (�$6,000) far lessthan the cost of commercially available Ar+ and Ti:Sapph laser systems (= $20,000) typically used inconfocal Raman systems.

30.3.1.2 Detectors and Spectrometers

A typical dispersive Raman detection system consists of a short focal length imaging spectrographattached to a cooled CCD camera. Clinical implementation of the Raman system requires spectralacquisition on the order of a few seconds. This fast acquisition in turn needs a fast spectrograph and ahighly sensitive detector, especially given the weak nature of the Raman signal. A typical CCD cameraused in spectroscopy consists of a rectangular chip wherein the horizontal axis corresponds to thewavelength axis. The vertical axis is used to stack multiple fibers for increased throughput and is subse-quently binned for improved SNR. Technological advances have led to CCD chips with quantum effi-ciencies on the order of 90% in the near-infrared. While various types of chips are commercially availablefor different applications, a back-illuminated, deep depletion CCD is highly recommended for NIR

©2003 CRC Press LLC

Page 7: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

Raman spectroscopy. These chips are known to be susceptible to the so-called “etaloning effect” whereinthe thin silicon chip acts as an etalon, resulting in the introduction of sharp peaks in the sample signalthat are hard to resolve from the narrow Raman signal. However, commercial cameras are now available(such as that offered by Roper Scientific Inc., Princeton, NJ) that incorporate technology that effectivelyeliminates this etaloning effect.

Most laboratory-grade systems currently utilize a liquid nitrogen–cooled CCD that allows the detectorto be cooled to at least –80�C, resulting in extremely low dark noise as well as read-out noise. While thislevel of noise is important in a laboratory grade system where integration times and other systemparameters are still being evaluated, a clinical system with a liquid nitrogen–cooled detector is cumber-some and impractical. Most recently, thermoelectrically cooled detectors are now available with multistagePeltier cooling systems that allow the camera to be cooled down to –80�C, thus solving this dilemma. Itshould be noted that shot noise due to tissue fluorescence continues to be a limiting factor.

Silicon-based detectors are relatively inexpensive and have high quantum efficiencies in the NIR,making them ideal detectors down to about 1100 nm. For longer wavelengths, other types of detectorssuch as indium gallium arsenide (InGaAs) and germanium (Ge) detectors need to be used. While thelonger wavelengths do result in lower fluorescence interference, the quantum efficiency and noise exhib-ited by these detectors are less than silicon detectors. Research needs to be conducted in this area to verifythat Raman spectra with adequate S/N can be acquired from tissue using 1064 nm excitation. Newtechnological advances have yielded low-cost InGaAs and Ge detectors with compact Nd:YAG lasersresulting in extremely compact 1064 nm Raman systems (TRI, Inc., Laramie, WY).

A Raman-sensitive detection system requires an appropriate imaging spectrograph that couples to thelight delivery device (such as a fiber probe) on one end and the CCD of choice at the other end. In orderto resolve details of the Raman bands, the Raman system should have a spectral resolution of at least 8cm–1. Commercial spectrographs optimized for Raman use are now available that provide f-numbermatching with standard fiber optics and high throughput for rapid acquisition using holographic gratings.These spectrographs (such as that offered by Kaiser Optical Systems, Inc., Ann Arbor, MI) are compactby design and rugged for portable use. Additional components of the detection system include laser linerejection filters that remove any laser light as well as the elastically scattered light from the detected signal.Holographic notch filters can block the laser wavelength with an optical density of six with steep edgesand provide 90% transmission elsewhere with a relatively flat curve.

Most Raman systems used in medical applications today are state-of-the-art, high-cost, high-perfor-mance Raman systems.27,35 This system at its best consists of a Kaiser imaging spectrograph, a RoperScientific liquid nitrogen–cooled CCD detector, a 300 mW Process Instruments diode laser, and a customfiber probe. Typical integration times are on the order of 5 to 10 sec for this type of system.

30.3.1.3 Light Delivery and Collection (Fiber Optics)

Raman scattering is a weak phenomenon, but most materials are Raman active, which allows for molec-ular-specific study of samples. On the other hand, because most molecules are Raman active, the materialsused in the Raman system themselves interfere with the detection of sample signal. The developmentand availability of diode lasers, imaging spectrographs, and cooled CCD cameras has made it possibleto build compact NIR Raman systems that acquire spectra with short integration times. However, thelimiting factor remains the signal generated in the delivery systems (luminescence and Raman) used forremote sensing.36 Light is typically delivered using optical fibers made of silica in a remote sensingspectroscopic system. However, silica has a strong Raman signal, which overrides sample signal. Thesignal generated is proportional to the fiber length and limits the detection capability of the technique.37

Figure 30.4 shows the Raman spectrum of fused silica fibers used to design a probe. Raman signal wasobserved to be generated from the core as well as the cladding and buffer of the fiber.20,37 This signal canhave magnitudes equal to and sometimes greater than those of the sample under study and thus demandscareful consideration. Fiber signal is generated in the delivery fiber by the excitation light. In addition,background signal is generated in the collection fibers by the elastically scattered excitation light returningto the collection fibers. Mathematical techniques cannot be used to remove this unwanted fiber signal.

©2003 CRC Press LLC

Page 8: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

A feasible probe design must prevent unwanted signal generated in the delivery fiber from illuminatingthe sample as well as prevent elastically scattered excitation light from entering the collection fibers andgenerating unwanted signal.

Several different designs have been proposed for potential clinical acquisition of Raman spectra usingfiber optic probes. Early on, Myrick and Angel developed different dual fiber probes that could be usedunder different conditions with maximum collection efficiency but minimum fiber interference.36 Mostfiber designs since have been based on similar concepts with modifications. In general, Raman probedesigns utilize a bandpass filter placed after the excitation fiber lens, thus allowing only the transmissionof the excitation light from the delivery fiber, and a longpass or notch filter placed in front of the collectionfibers that blocks the transmission of the Fresnel reflected excitation light and also prevents the elasticallyscattered light from entering the collection fibers. These filters are placed at the tip of the probe formaximum effectiveness and they must be sized on the order of a few millimeters or smaller. There is ademand in biomedicine for high-quality optical coatings and micro-optical components to simplify thedesign of much-needed compact fiber optic probes for Raman spectroscopy.

Multiple fiber bundles have been utilized for fluorescence measurements of tissue by several groups.6–8

Such a bundle typically consists of a central excitation fiber surrounded by many collection fibers linearlyaligned in front of the spectrometer. McCreery et al. used this design and tested it on different samples.38

Although spectra with good SNR ratio could be obtained from transparent samples, fiber backgroundwas still a serious problem in samples with high elastic scattering such as tissue. Feld et al. have adaptedan old design (used in solar energy collection) for improved signal collection to allow spectral acquisitionin a few seconds.39 A hollow compound parabolic concentrator (CPC) was used at the distal tip of theprobe to yield signals with seven times greater collection than a fiber probe without the CPC. Fiberbackground was reduced by using a dichroic mirror and separate excitation and collection fiber geom-etries. Excitation light was reflected by the mirror (placed at 45� with respect to the sample normal)through the CPC onto the sample. Raman spectra were collected by the CPC, transmitted through themirror into a collection fiber bundle (with about 100 fibers) to a detector. This probe design was usedto acquire Raman spectra for transcutaneous blood glucose measurements.

In vivo applications thus far have been mostly confined to exposed tissue areas where fiber backgroundcould be circumvented using a macroscopic arrangement — e.g., noncontact probe (DLT, Laramie, WY)used for breast tissues by McCreery et al.32 and CPC probe for transcutaneous measurements of bloodanalytes by Feld et al.42 However, other applications such as in the colon, cervix, and oral cavity, require

FIGURE 30.4 Raman spectrum at 785-nm excitation of fused silica fiber. (Modified from Berger, A.J. et al., Appl.Opt., 38, 2916, 1999.)

0.00E + 00

1.00E + 03

2.00E + 03

3.00E + 03

4.00E + 03

5.00E + 03

6.00E + 03

7.00E + 03

8.00E + 03

300 500 700 900 1100 1300 1500 1700

Raman Shift (cm−1)

Inte

nsi

ty (

a.u

.)

©2003 CRC Press LLC

Page 9: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

a more compact configuration and probe design. Thus, while many reports have been published assessingthe feasibility of Raman spectroscopy for disease detection, few reports have applied this technique invivo. The main obstacle continues to be fiber probe design.

One of the first designs of a compact fiber probe used clinically in the cervix was reported by theauthor.40 As stated previously, a feasible probe design must prevent unwanted signal generated in thedelivery fiber from illuminating the sample, as well as prevent elastically scattered excitation light fromentering the collection fibers and generating unwanted signal. Experimental results show that significantproportions of silica signal are generated in the excitation and collection fibers, indicating the need forfilters in the excitation and collection legs of the probe. Figure 30.5 shows a transverse section of thedesign that was implemented as a Raman probe for the cervix.40 Rather than placing the excitation legat an angle, a mirror surface is placed in the excitation path to deflect the beam onto the sample.

This probe was designed using the smallest available physical dimensions of the bandpass and theholographic notch filters at that time. An angularly polished gold wire was glued in place such that thedeflected excitation and normal collection spots overlap. One of the collection fibers was used to providean aiming beam during placement of the probe on the sample. A quartz shield was used at the tip of thecommon end of the probe, forming a barrier between the probe optics and the sample. Quartz wasselected as the material of choice because its fluorescence and Raman signal were known and anyadditional background signal from the probe could be identified. The inner surfaces of the metal tubingsused to house the probe optics were anodized to reduce the incidence of multiple reflections of light.

The probe was tested in a pilot clinical study approved by the Institutional Review Board (IRB). NIRRaman spectra of the cervix were successfully measured in vivo, as shown in Figure 30.6, using integrationtimes in the order of 90 sec.43 Since this initial report was produced, several designs for a Raman probehave been reported. One design in widespread use is the beam-steered Enviva Raman probe designed byVisionex Inc. (now Cirrex Corp., Atlanta, GA), which incorporates in-line filtering of the laser light andthe scattered light at the tip of the probe.41 The beam steering allows improved overlap of the excitationand collection volumes and thus improves signal collection efficiency (see Figure 30.3).

Another unique feature of this probe was its dimension in the order of 1 to 3 mm, which was not seenin other probes available commercially. However, the company no longer manufactures these probes and

FIGURE 30.5 Transverse section of the clinical Raman probe implemented in the cervix.

WindowSample

8 mm 4 mm

50, 100 µmCollection Fibers

Single 200 µmExcitation Fiber

HolographicNotch Filter

BandpassFilter

Gold Wire

©2003 CRC Press LLC

Page 10: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

thus a need for such slickly designed Raman probes still exists. Attempts have been made by othercompanies such as InPhotonics Inc. to fill the gap. In a survey of currently available probes for biomedicalRaman use, it was found that all commercially available probes were bulky (up to several millimeters indiameter) and expensive (�$3500 or more). Proponents of Raman spectroscopy for tissue diagnosis mustrely on custom-designed fiber probes built using commercial vendors.

Other new designs have been recently reported. Most recently, the Feld group has proposed a newerdesign for breast cancer detection, based on micro-optics and optical coatings.42 Results of spectraacquired with this probe have not yet been reported in vivo. The recent developments in the field ofmicrofabrication (MEMs) open a new avenue for compact probe designs using reflective optics. Others(including the author) continue to develop new ideas for feasible probe designs that should be testedand reported in the near future.

30.4 Effect of Sample Conditions

The design and construction of a Raman system is clearly a critical part of the process. An often forgottenaspect in the use of a new technique deals with the application itself. In developing and applying a newtechnique, the technology must first be tested in a model system before it can be applied in vivo.Subsequently, when it is ready for in vivo use, several factors unique to clinical use must also be considered.These issues are addressed below.

30.4.1 Tissue Model Selection

In the field of spectroscopy, there are two philosophies on the approach to use in testing the feasibilityof a given technique. Some researchers prefer to use an animal model that resembles human tissues infunction and structure to simulate the behavior of human tissues, while others prefer in vitro humantissue studies before tackling in vivo studies. Each of these approaches needs to be carefully consideredin the context of the human in vivo model.

Although an animal model allows in vivo testing, Raman spectra of animal tissues can in some casesdiffer from those of human tissues, resulting in conclusions not valid for humans. One example of thisis the use of mouse eye lenses for the study of lens aging.43 In early studies, Yu suggested that the intensityof Raman bands at 2580 cm–1 due to sulfhydryl groups (–SH) and 508 cm–1 due to disulfide groups(–S–S) can be used to calculate their relative concentrations. These can then be related to aging and

FIGURE 30.6 In vivo NIR Raman spectra of cervical tissue measured using the probe shown in Figure 30.5.

0

10

20

30

40

50

600 800 1000 1200 1400 1600 1800

Raman Shift (cm−1)

Inte

nsi

ty

Normal

HG SIL

936

1066

13211452

1659

©2003 CRC Press LLC

Page 11: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

cataract formation. In an extensive study of mouse lenses, a fall in –SH concentration and a correspond-ing increase in –S–S concentration was observed along the visual axis of mouse lens nucleus from age1 to 6 months. A tandem increase in protein development with a decrease in sulfhydryl was concludedto accompany the normal aging process. However, on repeating these studies on guinea pig and humanlenses, a different phenomenon was observed. Although guinea pig lenses also showed a decrease insulfhydryl intensity, no corresponding increase in disulfide band was observed. Human eye lensesbehaved similarly to guinea pig lenses. This study clearly indicates the pitfalls in improper selection ofan animal model and that independent verification of the model is needed to ensure that results will beapplicable to human conditions.

A primary concern in devising a clinical diagnostic system based on Raman spectroscopy is whetherthe spectra of excised tissue resemble those of tissue spectra acquired in vivo and whether the informationobtained from these in vitro studies can be applied to a clinical setting. It is well known that the spectraacquired from in vitro tissues (especially those that have undergone the freeze–thaw process) differ fromthose acquired in vivo.44 Spectra from intact human stratum corneum were compared to those fromexcised human stratum corneum;45 significant differences in the Raman spectral features were observed.Increased intensity of the C–C stretching vibrational bands at 1030 and 1130 cm–1 was observed in vivo.An additional spectral band at 3230 cm–1 of unknown origin is observed only in vivo. Thus, it is importantto be aware of the potential differences that may occur when moving from in vitro studies to in vivoconditions. Most researchers who follow this approach use the in vitro studies as proof of concept beforemoving directly into in vivo studies.

30.4.2 In Vivo Considerations

In vivo clinical application of the technology is the goal of all researchers in this field. However, variousissues specific to working in vivo must be considered in the planning of these clinical studies. Logisticissues such as IRB approval and consent forms have become challenging processes. When building aclinical system, some issues need to be considered. Because the probe must be sterilized or disinfected,depending on the application, it must be designed to withstand the physical and chemical conditions ofthese processes. The power consumption of the Raman system must not overload hospital and clinicsystems. Such requirements can be verified by the biomedical engineering department of the hospital.Acquisition times must be on the order of a few seconds for the sake of both the patient and clinicianparticipation. The system needs to be self-contained and nonobtrusive. Some of these issues may appeartrivial, but they all contribute to smooth conduct of clinical studies.

Any technique that uses lasers must comply with certain safety standards to avoid potential hazards.Safety standards for laser exposure of skin and eye are established by the American National StandardsInstitute (ANSI). The power densities used for successful Raman spectral acquisitions can be quite highand thus warrant consideration as safety hazards. The maximum permissible exposures as set by ANSIare wavelength dependent and can be calculated for a given exposure time following the directionsspecified in the ANSI laser safety manual.46

Sources of variability that need to be evaluated experimentally include ambient effects, physicianvariation, pressure effects, and procedural interference. Regardless of the short duration of the integrationtimes used for Raman acquisition, ambient effects can be significant for spectral integrity. Although roomlights may be turned off and windows may be light-tight in a laboratory setting, these are more criticalissues in a clinical setting and require careful planning. Variability from physician to physician in termsof single-site placement, as well as site-to-site placement for a given pathology, must be evaluated. Theeffect of varying pressures applied to the probe during acquisition should also be considered. Anothercritical parameter is the measurement-site to biopsy-site variability. Because histology continues to bethe gold standard against which we must compare this technology, it is critical that the biopsy be obtainedfrom the precise site of spectral measurement using such techniques as marking inks, etc. placed appro-priately. It is essential to account for sources of variability when acquiring Raman spectra from humanpatients in vivo.

©2003 CRC Press LLC

Page 12: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

30.5 Processing

Once a Raman spectrum is acquired using a feasible Raman system, it must be processed for differentreasons, including calibration, noise smoothing, and fluorescence elimination and binning. Other pro-cessing methods may be applied depending on the requirements of the analysis tools and the variabilityin the data.

30.5.1 Calibration

Because different researchers have different approaches in the development of a given application forRaman spectroscopy, a standard must be set to allow transferability of data across systems and methods.Instrument and data calibration are key components of data processing immediately following dataacquisition. Some calibration must occur at the time of acquisition and other calibration may be imple-mented later. Most biomedical Raman systems currently in use are based on dispersive CCD systemswhere the spectra are recorded as a function of pixel numbers along the horizontal axis. Thus, thespectrum of a known calibrated source such as a neon lamp is typically taken to calibrate the horizontalaxis. Raman spectrum of a strong Raman scatterer such as naphthalene, a strong Raman scatterer andfluorophore such as rhodamine 6G, and a weak Raman scatterer such as methylene blue are used asintensity standards to account for day-to-day variations in the spectral intensity. A spectrum of the lasersource is taken to verify the location (and thus the wavelength) of the laser line. This spectrum is alsoused with the neon spectrum to convert the horizontal axis in terms of relative wavenumbers.

Finally, there exists a wavelength-dependent efficiency of all the various components of the detectionsystem such as the grating, the filters, optics, quantum efficiency of the CCD chip, etc.; therefore, thewavelength dependency of the system response must be calibrated. This is typically done using an NIST-calibrated source such as a tungsten lamp to generate correction factors for the instrument variation.67

As various researchers acquire calibrated Raman spectra, a library of tissue Raman spectra is createdthat can be used to study cross links in the chemical compositions and correlation in tissue pathologiesas well as to strengthen and standardize the use of Raman spectroscopy for tissue diagnosis.

30.5.2 Fluorescence Elimination

In recent years we have witnessed an explosion in the use of Raman spectroscopy for biological purposessuch as tissue diagnosis, blood analyte detection, and cellular examination. The greatest benefits of thistechnique are its high sensitivity to subtle molecular (biochemical) change and its capacity for nonin-trusive application. Because biological applications of Raman spectroscopy involve turbid, chemicallycomplex, and widely varying target sites, much of the challenge in using Raman spectroscopy for bio-logical purposes is not only the acquisition of viable Raman signatures but also the suppression of inherentnoise sources present in the target media. Perhaps the greatest contributor of “noise” to biological Ramanspectra is the intrinsic fluorescence of many organic molecules in biological materials. This fluorescenceis often several orders of magnitude more intense than the weak chemical transitions probed by Ramanspectroscopy and, if left untreated, can dominate the Raman spectra and make analysis of tissue bio-chemistry as probed by the technique impractical. Therefore, in order to extract Raman signal from theraw spectra acquired, it is necessary to process the spectrum to remove this fluorescence.

A number of techniques, implemented in both hardware and software, have been proposed for fluo-rescence subtraction from raw Raman signals. Hardware methods such as wavelength shifting and time-gating have been shown to effectively minimize fluorescence interference in Raman spectra, but requiremodification of the spectroscopic system to achieve their results.48,49 Mathematical methods implementedin software require no such system modifications, and have thus become the norm for fluorescencereduction. These methods include first- and second-order differentiation,50,51 frequency-domain filter-ing,42 wavelet transformation52,53 and manual polynomial fitting.20, 54 Although each of these methods hasbeen shown to be useful in certain situations, these methods are not without limitations.

©2003 CRC Press LLC

Page 13: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

Differentiation can be implemented in various ways. One way is to measure the spectra at two slightlyshifted excitation wavelengths and take their difference.55 The fluorescence remains unchanged at bothexcitation wavelengths whereas the Raman peaks are shifted. The difference of the two spectra is com-parable to the first derivative of the Raman spectrum; integrating the difference spectrum yields theoriginal Raman signal. A similar result can be obtained by measuring the spectrum at a single excitationwavelength and taking the first derivative of the spectrum. The Raman spectrum can be obtained byintegrating the noise-smoothed derivative spectrum following baseline correction.49,50 The derivativemethod is entirely unbiased and very efficient in fluorescence subtraction, yet it severely distorts Ramanline shapes and relies on complex mathematical fitting algorithms to reproduce a traditional spectralform.49

Frequency filtering can be achieved by using the fast Fourier transform (FFT). In this technique, themeasured spectrum is Fourier transformed to the frequency domain by taking the FFT of the signal. TheFFT signal can then be multiplied with a linear digital filter to eliminate the fluorescence. The inverseFFT then yields the Raman spectrum free of fluorescence.52 This method can cause artifacts to begenerated in the processed spectra if the frequency elements of the Raman and noise features are notwell separated. Wavelet transformation, a more recently utilized method, is highly dependent on thedecomposition method used and the shape of the fluorescence background.

A simple, accurate, and more elegant method to subtract fluorescence is to fit the measured spectrumcontaining both Raman and fluorescence information to a polynomial of high enough order to describethe fluorescence line shape but not the higher-frequency Raman line shape (Figure 30.7). A fourth- orfifth-degree polynomial has been found to be optimal.20 Polynomial curve-fitting has a distinct advantageover other fluorescence reduction techniques because of its ability to retain the spectral contours andintensities of the input Raman spectra. However, simply fitting a polynomial curve to the raw Ramanspectrum in a least-squares manner will not efficiently reproduce the fluorescence background becausethe fit is based on minimizing differences between the fit and the measured, which includes the fluores-cence background and the Raman peaks. Subsequent subtraction of this fit polynomial results in aspectrum that varies about the zero baseline.

This technique traditionally relies on user-selected spectral locations on which to base the fit fromregions that do not include the more intense Raman bands. Unfortunately, this intervention has severaldrawbacks. It is time consuming because the user must process each spectrum individually to identify

FIGURE 30.7 Processing Raman spectra of rhodamine 6G for fluorescence removal using a manual fifth orderpolynomial fit.

0

5

10

15

20

25

30

35

600 800 1000 1200 1400 1600 1800

Raman Shift (cm−1)

Inte

nsi

ty

Processed Raman spectrumFluorescence background Measured glucose spectrum

X3

©2003 CRC Press LLC

Page 14: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

non-Raman-active spectral regions to be used in the fit. In addition, identification of non-Raman-activefrequencies is not always trivial because biological Raman spectra sometimes contain several adjacentpeaks or peaks that are not immediately obvious. The end effect is a method that is highly subjectiveand prone to variability. In order to address the limitations of existing methods of fluorescence subtrac-tion, various methods to automate polynomial curve-fitting have been developed that retain the benefitsof manual curve fitting, without the need for user intervention.56 One implementation of this automationis the modified-mean method.56

In one implementation of this automation, the basis is a simple sliding-window mean filter, in whichthe center pixel intensity in the window is set equal to the mean of all the intensities within the window.The modification involved in this method is that any filtered pixels with an intensity value higher thanthe original are automatically reassigned to the original intensity; therefore, the smoothing only eliminateshigh-frequency Raman peaks while retaining the underlying shape of the baseline fluorescence. Thissmoothing is repeated over the entire spectrum until a specified coefficient of determination (R2) of theprocessed spectrum with an nth order polynomial (typically fifth order) is obtained. The smoothedspectrum is then subtracted from the raw spectrum to yield the Raman bands on a near-null baseline.This algorithm can be implemented easily in MATLAB, thus automating the entire procedure.

Figure 30.8 shows the Raman spectra of rhodamine 6G, as well as the processed Raman spectrumusing derivative, FFT, and automated polynomial fluorescence subtraction techniques for direct compar-ison of the methods. The figure shows the effectiveness of the automated polynomial method for removingthe slow varying fluorescence baseline while retaining the Raman features of rhodamine, especially incomparison to the other methods.

Each of the different techniques has advantages and disadvantages and the method should be selectedbased on the specific application and measurement technique used. Mosier-Boss et al. tested the use ofthe shifted excitation, first derivative, and FFT techniques for fluorescence subtraction; a preference forusing the FFT technique was indicated due to its ability to filter out random noise from the spectrum.49

In an analysis of the different techniques, the use of a polynomial fit was found to be the simplesttechnique from both experimental and computational points of view.

FIGURE 30.8 Raman spectra of rhodamine 6G processed using the derivative, manual curve-fitting, and modified-mean methods.

400 600 800 1000 1200 1400 1600 1800

0

0

0

Raman Shift (cm−1)

Inte

nsi

ty (

a.u

.)

RawSignal

Derivative

Automated Polynomial

Fourier

©2003 CRC Press LLC

Page 15: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

30.5.3 Other Preprocessing Methods

Because Raman scattering is such a weak phenomenon, the SNRs of most measured Raman spectra aresuch that noise-smoothing filters must be included in the processing procedure. Various types of filtersthat have been effectively used include the Gaussian filter, whose full-width-half-maximum is typicallyset equal to the spectral resolution of the system, and the Savitsky-Golay filter of various orders.27 Whenusing any method of noise smoothing, care should be taken to retain the integrity of the spectral lineshape, especially in the case of multiple peaks.

Other preprocessing methods include binning of the spectral dataset for computational ease becauseof a large number of variables. Depending on the variability in the acquired data and the needs of theanalysis methods used, other methods include normalization to intensity standards, normalization to itsown maximum intensity, mean-scaling the spectra acquired from a given patient, etc.

30.6 Analysis

One advantage of spectroscopic diagnosis is automation, which allows objective and real-time diagnosisof pathologies. Differences in spectral features can be incorporated into diagnostic algorithms; severaltechniques have been identified and applied to enhance the differentiation and classification of tissuesfor potential automated clinical diagnosis. Because Raman spectroscopy is a molecular-specific technique,the contribution of the various participating chromophores can also be extracted from the measuredspectra. These can, in turn, be used in diagnostic algorithms and in the understanding of the spectralsignature as it pertains to the disease process. In addition, concentration of specific components such asglucose can be obtained for diagnostic use.

30.6.1 Automated Diagnosis

Early studies analyzed biological Raman spectra for differences in intensity, shape, and location of thevarious Raman bands between normal and non-normal materials such as tissues. Based on consistentdifferences observed among the various tissue categories, diagnostic algorithms are developed usingempirical methods. These algorithms may be based on changes in intensity or ratios of intensities ornumber and location of peaks. For example, it has been observed that the intensity ratio of the CH2

bending vibrational mode at 1450 cm–1 to the amide I vibrational mode at 1655 cm–1 varies with diseasein several applications, including breast, gynecologic, and precancers.57 These empirical algorithms indi-cate the specific changes that occur in the spectra acquired and provide information about the biochemicalprocesses that result in these spectral differences.

This empirical analysis, however, has two important limitations. First, clinically useful diagnosticinformation is typically contained in more than just the few wavenumbers surrounding peaks or valleysobserved in tissue; a method of analysis and classification that includes all the available spectral infor-mation can potentially improve the accuracy of detection. Second, empirical algorithms are optimizedfor the spectra within the study. Hence, the estimates of algorithm performances will be biased towardthat tissue population. An unbiased estimate of the performance of the algorithms is required for anaccurate evaluation of the performance of Raman spectroscopy for tissue diagnosis. To address theselimitations, multivariate statistical techniques have become the practice to develop and evaluate algo-rithms that differentiate between normal and non-normal tissues.

In recent years, the potential for using multivariate techniques for spectroscopic data analysis indisease detection has been exploited with great success.20 Discrimination techniques such as linearregression, as well as classification techniques such as neural networks, have been used.20,22 Datacompression tools such as principal component analysis (PCA) are used to account for variability inthe data.58 Subsequently, methods such as hierarchical cluster analysis (HCA),11 linear discriminantanalysis (LDA)22 and others have been used to yield classification algorithms for disease differentiation.Partial least squares, a regression-based technique, and hybrid linear analysis have been used to extract

©2003 CRC Press LLC

Page 16: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

accurate concentrations of analytes such as glucose using NIR Raman spectra for transcutaneous bloodanalysis.59

The consensus of the scientific community is that, while any of these methods may form the basis ofperformance estimates, the true measure of success of Raman spectroscopy for tissue diagnosis can beassessed by extensive unbiased (and independent) validation alone. This may best be accomplished byrandom distribution of the subject population into two equal datasets: one for calibration or testing andthe other for validation. When this is not feasible, one can then rely (in a limited manner) on the leave-one-out or leave-five-out methods of cross validation.

30.6.2 Component Extraction

Analyzing Raman spectra for the purpose of disease classification may be sufficient when the goal of thestudy is to achieve tissue diagnosis alone; however, most researchers want to know if and why they candiagnose disease. In the case of Raman spectroscopy, the wealth of information provided in these spectra,especially with respect to biochemical composition, allows us to study the question of “why” in detailthat can subsequently be used for the purpose of detection as well.

While most researchers study the biochemical basis of the measured Raman spectra from biologicaltissues, some groups have been particularly active in this area. For example, Puppels et al. have utilizedskin Raman spectra to obtain quantitative information about skin hydration;13 Feld et al. have developedcomprehensive models for biochemical component extraction that have subsequently been used for diseaseclassification.60 Each of these methods is based on the acquisition of Raman spectra from individuallyidentified chromophores as morphological tissue components or as extracted biochemical constituents.

Pixelated Raman microspectroscopy is typically used (with or without confocality) to measure Ramanspectra from individual morphologic tissue components using tissue sections. A Raman spectrometer iscoupled to a microscope and is scanned across the tissue section to obtain Raman images that can thenbe correlated with serial hematoxylin- and eosin-stained sections to identify relevant morphologic com-ponents and their Raman signature. Alternatively, tissue chromophores can be extracted from biochemicalassays and Raman spectra can be acquired from each of these extracted chromophores.60 By usingmathematical models such as those developed by the Feld group, contributions of the various extractedor morphologic components to intact tissue spectra can then be extracted.60

30.7 Clinical Application of Raman Spectroscopy

In vivo studies are necessary to fully evaluate the potential of Raman spectroscopy for clinical diagnosis.Several groups have initiated this process with varying degrees of success.26,27,62 This progression has beenmade possible by the development of sensitive instrumentation, use of fiber optics, and development ofautomated algorithms. However, despite these significant developments, the majority of reported studieson the application of Raman spectroscopy for tissue diagnosis continue to be from in vitro studies. Onlya limited number of reports exist on the application of Raman spectroscopy in vivo in humans in aclinical setting; most of them are pilot clinical studies. Nevertheless, these critical studies create the roadmap for the future of Raman spectroscopy in biomedicine.

Several biological molecules such as nucleic acids, proteins, and lipids have distinctive Raman featuresthat yield structural and environmental information. These molecules have been studied in solutionsand in their natural microscopic environments.20 The molecular and cellular changes that occur withdisease result in distinct Raman spectra that can be used for diagnosis. The transitional changes inprecancerous tissues as well as in benign abnormalities such as inflammation can also yield characteristicRaman features that allow their differentiation.

Several groups have indicated the potential of vibrational spectroscopy for disease diagnosis in variousorgan sites. These groups have shown that features of the vibrational spectrum can be related to molecularand structural changes associated with disease. Raman spectroscopy has been studied extensively fortissue diagnosis in four main organ sites: breast,60 esophagus (and the GI tract),63 cervix (and other

©2003 CRC Press LLC

Page 17: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

gynecological tissues),27 and skin.64 Other organ sites include the brain, the eye, and biological fluids.11,20,21

Some examples of the application of Raman spectroscopy for disease diagnosis are detailed in thefollowing sections.

30.7.1 Breast

Until recently, perhaps the most widespread application of Raman spectroscopy in cancer research hasbeen for breast cancer detection. Breast cancer is the most common cancer in women, accounting for18% of all cancer deaths in women.65 The breast consists of mammary glands arranged in lobes separatedby fibrous connective tissue and a considerable amount of fatty tissue.66 Fibrocystic changes are benignproliferative processes that vary from the innocuous to those associated with increased risk of carcinomas.These benign changes take three forms: (1) cyst formation and fibrosis, (2) hyperplasia, and (3) adenosis.In most afflicted breasts, all forms occur simultaneously in fibroadipose stroma.

Breast tumors may arise from the ductal and lobular epithelium or connective tissue. These tumorsvary from benign adenomas and malignant adenocarcinomas to benign fibromas and malignant fibro-sarcomas. Infiltrating ductal carcinoma (IDC), the most frequent invasive form of breast cancer, has beenstudied using Raman spectroscopy. IDCs typically show an increase in dense fibrous stromal tissue andmalignant proliferation of ductal epithelium.67 Although routine screening using mammography can aidin early detection of malignancy, lesions identified with this method must be biopsied and evaluatedhistopathologically to determine the presence of malignancy. In recent years, spectroscopic techniqueshave been used for breast cancer diagnosis.31,68 Although fluorescence spectroscopy has shown somepromise as a diagnostic tool, Raman spectroscopy may provide more definitive characteristics that allowdifferentiation of benign and malignant tumors.

Several groups have studied the potential of Raman spectroscopy for pathologies of the breast, fromdetection of breast cancers to study of capsules from breast implants.69 Using an FT-Raman system, Alfanoet al. obtained the first Raman spectra from excised normal human breast tissues and benign andmalignant breast tumors and discussed the feasibility of using FT-Raman spectroscopy for differentiatingnormal and malignant breast tissues.32 The vibrational spectra of benign breast tissues (which includenormal tissues) showed four characteristic bands at 1078, 1300, 1445, and 1651 cm–1. The spectra ofbenign tumors showed bands at 1240, 1445, and 1659 cm–1, and malignant tumors displayed only twoRaman bands at 1445 and 1651 cm–1. The intensity ratio of 1445 to 1651 cm–1 was found to be larger inbenign tissues compared to benign tumors and the same ratio was found to be even lower in malignanttumors.

Redd et al.30, 31 employed Raman spectroscopy using visible excitation to study excised human breasttissues. Spectra were also obtained from pure compounds, and the features observed in tissue spectrawere determined to be primarily due to carotenoids, myoglobin, and lipids. The Raman spectra of normalbreast tissues showed peaks assigned to carotenoids at 1005, 1157, and 1523 cm–1 (not observed with IRexcitation); peaks assigned to lipids, primarily oleic acid derivatives at 1302, 1442, and 1653 cm–1; a peakat 1370 cm–1 due to myoglobin; and several other smaller unassigned peaks.

A subsequent study by McCreery et al. assessed the feasibility of using NIR Raman spectroscopy forbreast cancer detection. Chromophore contributions differed as excitation was shifted from the visible(yielding carotenoid and lipid bands) to the NIR (yielding only lipid bands) in normal tissues. NIRRaman spectroscopy yielded signals with lower fluorescence interference and higher SNR.

NIR Raman spectra were measured from excised normal human breast tissues, tissues with fibrocysticchange, and infiltrating ductal carcinoma (IDC) at 784 nm excitation (Figure 30.9).31 The ratio of theareas under the peaks at 1654 cm–1 and 1439 cm–1 were found to increase in malignant breast tissues ascompared to normal breast tissues. This increase is consistent with the changes reported by Alfano et al.32

The intensity of the 1654 cm–1 C = C stretching band varies with the degree of fatty acid unsaturation;the CH2 scissoring band at 1439 cm–1 depends on the lipid-to-protein ratio. The spectra from IDC tissuesshowed an overall decrease in intensity with respect to normal tissue.

©2003 CRC Press LLC

Page 18: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

Several differences were also observed when IDC tissues were compared with benign abnormal tissue.In benign tissue, the intensities of the bands at 1656 cm–1 and 1259 cm–1 were smaller than the band at1449 cm–1 and this band is further shifted to 1446 cm–1. The region of 850 to 950 cm–1 showed only twobands in benign tissue as compared to four in IDC samples. The peaks observed in normal tissues wereprimarily attributed to oleic acid methyl ester, a lipid, and the peaks observed in IDC and benign tissueswere primarily attributed to collagen I. This is consistent with histopathology where IDC and benigntissues show an increase in interstitial tissues microscopically. These spectral differences were found tobe significant; subsequently, this group reported the development of a clinical probe and system for invivo testing. Preliminary testing of the probe was reported on in vitro samples but no other reports havesince been published.

Similar results were also obtained by Manoharan et al. using a comparable system.68 The Raman spectraacquired from this study were analyzed using multivariate statistical techniques and results yielded nofalse negatives. This group has subsequently reported the development of a clinical Raman system andin vivo testing will be reported in the future.42

30.7.2 Cervix

The various gynecologic tissues differ both structurally and functionally67,70 from one another and theyhave been studied using Raman spectroscopy. Because of its well-characterized disease process and itssignificance, the cervix has been studied extensively in the development of spectroscopic detection. Thecervix is the most inferior portion of the uterus, which typically measures 2.5 to 3.0 cm in diameter in

FIGURE 30.9 Raman spectra of (A) normal, (B) malignant and (C) benign human breast tissue at 784-nm excita-tion. (Modified from Frank, C.J., McCreery, R.L., and Redd, D.C., Anal. Chem., 67, 777, 1995.)

1260 1655

14491043

B

C

600 900 1200 1500 1800

1656

1446

1259

1043

Raman Shift (cm−1)

1656

1439

A

©2003 CRC Press LLC

Page 19: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

the human adult female. The cervix is covered by two types of epithelia. The multilayered squamousepithelium covers most of the ectocervix and is separated from the stroma by the basal layer; the columnarepithelium consists of a single layer of columnar cells and covers the surface of the endocervical canal.The interface of the two epithelia is called the squamo–columnar junction. Over time, the columnarepithelium is replaced by squamous epithelium, which causes the squamo–columnar junction to movetoward the os. This transitional epithelium is termed squamous metaplasia.70

Virtually all squamous cervical neoplasias (new growth) begin at the functional squamo–columnarjunction and the extent and limit of their precursors coincide with the distribution of the transformationzone.70 Cervical intraepithelial neoplasia (CIN) refers to the precancerous stages of cervical carcinomaand is often also referred to as cervical dysplasia. Here we generally refer to them as “precancers.” Otherpathologies that affect the cervix include cervicitis or inflammation, which is usually the response oftissue to injury,70 and the human papilloma viral (HPV) infection. Similarities observed in the mor-phological changes of the epithelial cells between those induced by HPV and precancer have led to thesuggestion that certain strains of HPV may be involved in the incipient stages of cervical precancer andother strains may aid in the progression of the disease.71 Thus, HPV is typically placed in the samecategory as mild precancers and is clinically treated as such. Endocervical cancers are typically adeno-carcinomas, arising within the endocervix as opposed to cervical epithelial lesions, which arise in thesquamo–columnar junction.70

Cervical cancer is the second most common malignancy found in women worldwide. It was estimatedthat in 2001, 4100 deaths occurred in the United States from this disease and 13,000 new cases of invasivecervical cancer were diagnosed.65 Although early detection of cervical precancer has played a central rolein reducing the mortality associated with this disease over the last 50 years, the incidence of preinvasivesquamous carcinoma of the cervix has risen dramatically, especially among women under the age of 50.The primary screening tool for cervical precancer is the Papanicolaou (Pap) smear, where scrapings fromthe walls of the ecto- and endocervix, which contain a variable number of cells, are examined anddiagnosed.70 Although the widespread application of the Pap smear as a screening tool has greatlydecreased the incidence of cervical cancer,72 sampling and reading errors lead to high false-positive and-negative rates. Treatment ultimately relies on directed biopsies and subsequent pathological findings.

Alfano et al. were the first to report on the feasibility of using FT-Raman spectroscopy for detectingcancers from various gynecologic tissues.57 Characteristic Raman eatures of normal tissues and malignanttumors from the cervix, uterus, endometrium, and ovary were described. Three significant peaks werenoted to differ in the Raman spectra of normal and benign cervix compared to cancerous lesions. Incancerous tissues, the intensity of the amide I stretching vibration band at 1657 cm–1 is less than theintensity of the C–H bending vibrational band at 1445 cm–1. The amide III band at 1262 cm–1 is broadenedin cancerous lesions. An additional unidentified peak at 934 cm–1 is observed only in normal and benigncervical samples. A possible diagnostic algorithm could be based on the relative intensities of the twopeaks where I1657 > I1445 in normal and benign tissues and I1657 < I1445 in cancerous samples. Alfano et al.attributed these peaks primarily to collagen and elastin.

The author has reported in vitro as well as in vivo studies on the application of Raman spectroscopyfor cervical precancer detection.27,58 In her early work, Raman spectra of cervical tissues were measuredto characterize the spectral signatures of the different cervical tissue types and to assess the feasibility ofusing Raman spectroscopy for cervical precancer diagnosis.58 Primary tissue Raman peaks were observedat 626, 818, 978, 1070, 1246, 1330, 1454, and 1656 cm–1 (�10 cm–1), present in all samples. The peaks at626, 818, and 1070 were attributed as primarily due to silica from the optics of the system. Both empiricaland multivariate techniques were used to explore the diagnostic capability of NIR Raman spectra fromcervical tissues. A multivariate discrimination algorithm developed using the entire Raman spectrumcould differentiate cervical precancers from nonprecancers with a sensitivity and specificity of 91 and90%, respectively.58 A discrimination algorithm developed using intensities at just 8 Raman bands gavean unbiased estimate of the sensitivity and specificity as 82 and 92%, respectively.58 Thus, the success ofthe in vitro Raman studies for precancerous cervical tissue recognition warranted the development of aclinical system.

©2003 CRC Press LLC

Page 20: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

In a pilot clinical study approved by the IRB, NIR Raman spectra of the cervix were successfullymeasured in vivo from 24 sites (11 normal, 4 inflammation, 4 metaplasia, 1 low-grade, and 4 high-gradelesions) in 13 patients.27 Raman spectra were measured from normal and abnormal areas of the cervixusing the fiber probe described in Figure 30.5 and a Raman system similar to the one shown in Figure 30.3.Figure 30.10 shows typical Raman spectra of the different types of cervical tissues acquired in vivo.27

Cervical tissue Raman spectra show peaks in the vicinity of 1070, 1180, 1210, 1245, 1270, 1330, 1400,1454, 1580, and 1656 cm–1. The ratio of intensities at 1454 to 1656 cm–1 is greater for precancerous tissuesthan for all other tissue types, while the ratio of intensities at 1330 to 1454 cm–1 is lower for sampleswith cervical precancer than for all other tissue types. A simple algorithm based on these two intensityratios separates high-grade lesions from all others with a sensitivity of 100% and a specificity of 95%.27

Preliminary results thus indicate that it is possible to measure Raman spectra in vivo and extract poten-tially diagnostically useful information.

In vivo Raman spectra, in general, appeared similar to in vitro Raman spectra previously obtainedfrom cervical biopsies. Due to the small number of patients included, the number of samples within aparticular category was small; only one of the investigated samples was histologically identified to be lowgrade. Thus, while promising results were obtained in discriminating high-grade lesions from all othertissue types, significant research still remains to be done to characterize the spectra of low-grade lesionsand other tissue pathologies of the cervix before Raman spectroscopy can be validated as a viable methodfor the diagnosis and potential screening of cervical precancers.73

30.7.3 Skin

The skin is the largest organ of the body and has many different functions. It is divided into two mainregions: the epidermis and the dermis. The epidermis primarily consists of keratinocyctes and varies inthickness throughout the body. Melanocytes, the pigment-producing cells of the skin, are found through-out the epidermis.66 The dermis consists mostly of fibroblasts, which are responsible for secreting thecollagen and elastin that give support and elasticity to the skin. There are two major groups of skincancers: malignant melanoma and nonmelanoma skin cancers. Cancers that develop from melanocytesare called melanoma. Nonmelanoma skin cancers are the most common cancers of the skin. Two of themost common nonmelanoma types are basal cell carcinoma (BCC) and squamous cell carcinoma (SCC).

FIGURE 30.10 In vivo NIR Raman spectra for different types of cervical tissues. (Modified from Utzinger, U. et al.,Appl. Spectrosc., 55, 955, 2001.)

Raman Shift (cm−1)

Ave

rag

e o

f p

eak

inte

nsi

ties

1000 1100 1200 1300 1400 1500 1600 1700 1800

Normal (11)

Inflammation (4)

Metaplasia (4)

High Grade (4)

©2003 CRC Press LLC

Page 21: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

About 75% of all skin cancers are BCCs, which have a high likelihood of recurrence after treatment,either at the same site or elsewhere.65 SCCs account for about 20% of all skin cancers; these carcinomasare more likely to invade tissues beneath the skin and distant parts of the body than are BCCs.

Although most people believe they are not at risk for skin cancer, cancers of the skin (includingmelanoma and nonmelanoma skin cancers) are the most common of all cancers and account for abouthalf of all cases in the United States.65 The American Cancer Society estimated approximately 9800 skincancer deaths for 2002, 7800 from melanoma and 2000 from other skin cancers. Melanomas account forabout 4% of skin cancer cases but cause about 77% of skin cancer deaths. The number of new cases ofmelanoma found in the United States is on the rise. The American Cancer Society predicted 53,600 newcases of melanoma in the United States in 2002. For most skin cancer patients (including melanoma andnonmelanoma skin cancer), early diagnosis and thorough treatment (i.e., complete resection) are criticalfor a favorable prognosis.

Current diagnostic methods for skin cancers rely on physical examination of the lesion in conjunctionwith skin biopsy. Suspicious areas are selected upon visual inspection by the clinician, after which thoselesions are partially or wholly biopsied for complete histological evaluation.65 The biopsy is then sectionedand stained for pathological investigation and diagnosis. This protocol for skin lesion diagnosis is acceptedas the gold standard, but it is subjective, invasive, and time consuming; hence, there is considerableinterest in developing a noninvasive diagnostic tool that can accurately detect skin lesions noninvasivelyin real time, especially in the early stages.

Previous studies on skin cancer detection using optical spectroscopy have been primarily limitedto fluorescence and diffuse reflectance spectroscopy and have produced limited success.74,75 The inter-ference of skin pigment and external agents such as creams and soaps has been the major limitationto the success of these techniques. Despite the relative ease of studying the skin, the first publishedreports of skin Raman spectra appeared only in the early 1990s. These early studies, however, focusedon characterizing skin components, and included research on skin hydration, skin aging, and the effectof UV radiation.11

Williams et al. utilized FT-Raman spectroscopy to examine a number of skin features and correlatethe Raman spectra with the biochemical agents responsible.76,77 Caspers et al. used confocal Ramanmicrospectroscopy to ascertain the in vivo Raman signal characteristics emanating from each layer ofnormal human skin. They showed that each layer of the skin contained Raman features that could behighly correlated with the protein and chemical content of each respective layer.62 A recent study has alsoshown that it is possible to extract carotenoid Raman spectra in vivo from various skin tissues.78 Gniadeckaet al. utilized FT-Raman spectroscopy to successfully differentiate BCC from normal, healthy skin in vitroin 16 patients.79 Although promising results were reported, FT-Raman spectroscopy is not a feasibletechnique for clinical diagnosis.

A recent study exploring the use of confocal Raman microspectroscopy for skin cancer detection wasalso reported.80 Results from two pilot studies performed on a novel bench top confocal Ramanmicrospectrometer were encouraging. In the first study, Raman spectra were acquired in vitro fromnormal and various types of malignant skin tissues obtained from human patients undergoing excision.80

Figure 30.11 shows the Raman spectra of normal skin, BCC, SCC, and melanoma tissues at a depth whereoptimal differences between the various tissue types were observed. Key differences in the spectra areobserved at several Raman bands including those seen at 860, 940, 1120, 1220 to 1340, and 1550 cm–1.

Another study assessed the capability of the confocal Raman system to acquire Raman spectra in vivoand the variability in the Raman spectra from various types of normal skin. Raman spectra were acquiredfrom the dorsal area of the hand from volunteers of various ethnicity and skin color (white, African-American, east Asian, and south Asian) at various depths of tissue. Subsequent statistical analysis of theRaman spectra at various depths revealed significant variations of Raman signal in the upper strata ofthe skin, presumably due to differences in melanin content. However, Raman spectra from locationsdeeper in the skin (�60 �m) showed much more similarity visually and statistically. These results indicatethat confocal Raman spectroscopy can limit the measurement volume to specified depths of the skin,

©2003 CRC Press LLC

Page 22: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

which may circumvent the problems in skin color differences traditionally associated with opticalmeasurement.

30.7.4 Other Applications

The preceding examples show the various stages of development in the implementation of Ramanspectroscopy for disease detection in vivo. Additional applications of this versatile technique not specif-ically reported here include the detection of various pathologies in the brain81 and the GI tract, especiallythe esophagus,63 ovary,56 and colon.82 Raman spectroscopy has also been used to determine variousanalytes such as those in the blood. For example, studies have been reported on the determination ofglucose concentration in diabetes patients.59 A recent study reported the transdermal measurement ofblood glucose using tissue modulation for signal extraction.83 Researchers are using Raman spectroscopyin clinical settings and results appear promising. Raman spectroscopy is poised to move to the next levelof implementation.

30.8 Future Perspectives

Recent technological developments have made clinical implementation of Raman spectroscopy feasible.Advances in diode laser technology and CCD detector technology have contributed tremendously to thisprocess. However, the need still exists for compact Raman probes and integrated systems that wouldmake this technique even more versatile. Currently, several integrated systems are available for theapplication of Raman spectroscopy in environmental and manufacturing processes. A similar advance-ment is required in the field of biomedicine. The introduction of micro-optics and microfabrication canfurther aid in the development of compact small-diameter probes.

This chapter presents a review of concepts, instrumentation, and sample applications of nonresonanceRaman spectroscopy for disease detection in vivo. The success of the technique has led to the developmentof feasible clinical systems that can measure Raman signals from tissue with short collection times. Manyof these systems have already undergone preliminary testing and several others are currently in the processof being tested. These studies clearly indicate that clinical application of Raman spectroscopy is imminentand may be expected to change the face of disease detection in the near future. While more extensive

FIGURE 30.11 NIR Raman spectra of normal and cancerous skin lesions acquired in vitro using a confocal Ramanmicrospectrometer.

600 800 1000 1200 1400 1600 1800

860 940 1120 15501220-1340

normal

BCC

SCC

melanoma

Raman Shift (rel. cm–1)

©2003 CRC Press LLC

Page 23: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

studies are still needed to form a true assessment of the capability of Raman spectroscopy for diseasedetection, preliminary results are extremely encouraging. What makes this technique so invaluable is notonly its use for disease classification, but also the possibility of evaluating the acquired spectra forbiochemical composition. This technique is a viable clinical tool and an important research tool forfurthering both the technology of patient care and the understanding of the disease process.

References

1. Stummer, W., Novotny, A., Stepp, H., Goetz, C., Bise, K., and Reulen, H.J., Fluorescence-guidedresection of glioblastoma multiforme by using 5-aminolevulinic acid-induced porphyrins: a pro-spective study in 52 consecutive patients, J. Neurosurg., 93, 1003, 2000.

2. Alfano, R.R., Pradhan, A., Tang, C.G., and Wahl, S.J., Optical spectroscopic diagnosis of cancer innormal and breast tissues, J. Opt. Soc. Am. B, 6, 1015, 1989.

3. Kennedy, T.C., Lam, S. and Hirsch, F.R., Review of recent advances in fluorescence bronchoscopyin early localization of central airway lung cancer, Oncologist, 6, 257, 2001.

4. Hung, J., Lam, S., LeRiche, J.C., and Palcic, B., Autofluorescence of normal and malignant bronchialtissue, Lasers Surg. Med., 11, 99, 1991.

5. Gillenwater, A., Jacob, R., Ganeshappa, R., Kemp, B., El-Naggar, A.K., Palmer, J.L., Clayman, G.,Mitchell, M.F., and Richards-Kortum, R., Noninvasive diagnosis of oral neoplasia based on fluo-rescence spectroscopy and native tissue autofluorescence, Arch. Otolaryngol. Head Neck Surg., 124,1251, 1998.

6. Ramanujam, N., Mitchell, M.F., Mahadevan, A., Thomsen, S., Malpica, A., Wright, T., Atkinson,N., and Richards-Kortum, R., Spectroscopic diagnosis of cervical intraepithelial neoplasia (CIN)in vivo using laser-induced fluorescence spectra at multiple excitation wavelengths, Lasers Surg.Med., 19, 63, 1996.

7. Cothren, R.M., Richards-Kortum, R.R., Sivak, M.V., Fitzmaurice, M., Rava, R.P., Boyce, G.A.,Hayes, G.B., Doxtader, M., Blackman, R., Ivanc, T., Feld, M.S., and Petras, R.E.., Gastrointestinaltissue diagnosis by laser induced fluorescence spectroscopy at endoscopy, Gastrointest. Endosc., 36,105, 1990.

8. Schomacker, K.T., Frisoli, J.K., Compton, C.C., Flotte, T.J., Richter, J.M., Nishioka, N.S., andDeutsch, T.F., Ultraviolet laser-induced fluorescence of colonic tissue: basic biology and diagnosticpotential, Lasers Surg. Med., 12, 63, 1992.

9. Johnson, C.R., Ludwig, M., O’Donnell, S., and Asher, S.A., UV resonance Raman spectroscopy ofthe aromatic amino acids and myoglobin, J. Am. Chem. Soc., 106, 5008, 1984.

10. Nie, S., Bergbauer, K.J., Ho, J.J., Kuck, J.F.R., Jr., and Yu, N.T., Applications of near-infrared Fouriertransform Raman spectroscopy in biology and medicine, Spectroscopy, 5, 24, 1990.

11. Choo-Smith, L.-P., Edwards, H.G.M., Endtz, H.P., Kros, J.M., Heule, F., Barr, H., Robinson, J.S.,Jr., Bruining, H.A., and Puppels, G.J., Medical applications of Raman spectroscopy: from proof ofprinciple to clinical implementation, Biopolymers, 67, 1, 2002.

12. Pilotto, S., Pacheco, M.T., Silveira, L., Jr., Villaverde, A.B., and Zangaro, R.A., Analysis of near-infrared Raman spectroscopy as a new technique for a transcutaneous non-invasive diagnosis ofblood components, Lasers Med. Sci., 16, 2, 2001.

13. Caspers, P.J., Lucassen, G.W., Bruining, H.A., and Puppels, G.J., Automated depth-scanning con-focal Raman microspectrometer for rapid in vivo determination of water concentration profiles inhuman skin, J. Raman Spectrosc., 31, 813, 2000.

14. Gellermann, W., Ermakov, I.V., Ermakova, M.R., McClane, R.W., Zhao, D.Y., and Bernstein, P.S.,In vivo resonant Raman measurement of macular carotenoid pigments in the young and the aginghuman retina, J. Opt. Soc. Am. A Opt. Image Sci. Vis., 19, 1172, 2002.

15. Colthup, N.B., Daly, L.H., and Wiberley, S.E., Introduction to Infrared and Raman Spectroscopy,Academic Press, Boston, 1990.

16. Ferraro, J.R. and Nakamoto, K., Introductory Raman Spectroscopy, Academic Press, Boston, 1994.

©2003 CRC Press LLC

Page 24: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

17. Carey, P.R., Biochemical Applications of Raman and Rresonance Raman Spectroscopies, AcademicPress, New York, 1982.

18. Twardowski, J. and Anzenbacher, P., Raman and IR Spectroscopy in Biology and Biochemistry, EllisHorwood, New York, 1994.

19. Feld, M.S. and Kramer, J.R., Mutagenicity and the XeCl excimer laser: a relationship of conse-quence? Am. Heart J., 122, 1803, 1991.

20. Mahadevan-Jansen, A. and Richards-Kortum, R., Raman spectroscopy for the detection of cancersand precancers, J. Biomed. Opt., 1, 31, 1996.

21. Hanlon, E.B., Manoharan, R., Koo, T.W., Shafer, K.E., Motz, J.T., Fitzmaurice, M., Kramer, J.R.,Itzkan, I., Dasari, R.R., and Feld, M.S., Prospects for in vivo Raman spectroscopy, Phys. Med. Biol.,45, R1, 2000.

22. Petrich, W., Mid-infrared and Raman spectroscopy for medical diagnostics, Appl. Spectrosc. Rev.,36, 181, 2001.

23. Mizuno, A. and Ozaki, Y., Aging and cataractous process of the lens detected by laser Ramanspectroscopy, Lens Eye Toxicol. Res., 8, 177, 1991.

24. Van de Poll, S.W., Romer, T.J., Puppels, G.J., and Van der Laarse, A., Raman spectroscopy ofatherosclerosis, J. Cardiovasc. Risk, 9, 255, 2002.

25. Buschman, H.P., Deinum, G., Motz, J.T., Fitzmaurice, M., Kramer, J.R., van der Laarse, A., Brus-chke, A.V., and Feld, M.S., Raman microspectroscopy of human coronary atherosclerosis: bio-chemical assessment of cellular and extracellular morphologic structures in situ, Cardiovasc. Pathol.,10, 69, 2001.

26. Shim, M.G., Song, L.M., Marcon, N.E., and Wilson, B.C., In vivo near-infrared Raman spectros-copy: demonstration of feasibility during clinical gastrointestinal endoscopy, Photochem. Photobiol.,72, 146, 2000.

27. Utzinger, U., Heintzelman, D.L., Mahadevan-Jansen, A., Malpica, A., Follen, M., and Richards-Kortum, R., Near infrared Raman spectroscopy for in vivo detection of cervical precancers, Appl.Spectrosc., 55, 955, 2001.

28. Yu, N.T. and East, E.J., Laser Raman spectroscopic studies of ocular lens and its isolated proteinfractions, J. Biol. Chem., 250, 2196, 1975.

29. Clarke, R.H., Hanlon, E.B., Isner, J.M., and Brody, H., Laser Raman spectroscopy of calcifiedatherosclerotic lesions in cardiovascular tissue, Appl. Opt., 26, 3175, 1987.

30. Redd, D.C.B., Feng, Z.C., Yue, K.T., and Gansler, T.S., Raman spectroscopic characterization ofhuman breast tissues: implications for breast cancer diagnosis, Appl. Spectrosc., 47, 787, 1993.

31. Frank, C.J., McCreery, R.L., and Redd, D.C., Raman spectroscopy of normal and diseased humanbreast tissues, Anal. Chem., 67, 777, 1995.

32. Alfano, R.R., Lui, C.H., Sha, W.L., Zhu, H.R., Akins, D.L., Cleary, J., Prudente, R., and Cellmer,E., Human breast tissues studied by IR Fourier transform Raman spectroscopy, Lasers Life Sci., 4,23, 1991.

33. Puppels, G.J., de Mul, F.F., Otto, C., Greve, J., Robert-Nicoud, M., Arndt-Jovin, D.J., and Jovin,T.M., Studying single living cells and chromosomes by confocal Raman microspectroscopy, Nature,347, 301, 1990.

34. Caspers, P.J., Lucassen, G.W., Wolthuis, R., Bruining, H.A., and Puppels, G.J., In vitro and in vivoRaman spectroscopy of human skin, Biospectroscopy, 4, S31, 1998.

35. Stone, N., Stavroulaki, P., Kendall, C., Birchall, M., and Barr, H., Raman spectroscopy for earlydetection of laryngeal malignancy: preliminary results, Laryngoscope, 110, 1756, 2000.

36. Myrick, M.L. and Angel, S.M., Elimination of background in fiber-optic Raman measurements,Appl. Spectrosc., 44, 565, 1990.

37. Myrick, M.L. and Angel, S.M., Comparison of some fiber optic configurations for measurementof luminescence and Raman scattering, Appl. Opt., 29, 1333, 1990.

38. Schwab, S.D. and McCreery, R.L., Versatile, efficient Raman sampling with fiber optics, Anal. Chem.,56, 2199, 1984.

©2003 CRC Press LLC

Page 25: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

39. Tanaka, K., Pacheco, M.T.T., Brennan, J.F., Itzkan, I., Berger, A.J., Dasari, R.R., and Feld, M.S.,Compound parabolic concentrator probe for efficient light collection in spectroscopy of biologicaltissue, Appl. Opt., 35, 758, 1996.

40. Mahadevan-Jansen, A., Mitchell, M.F., Ramanujam, N., Utzinger, U., and Richards-Kortum, R.,Development of a fiber optic probe to measure NIR Raman spectra of cervical tissue in vivo,Photochem. Photobiol., 68, 427, 1998.

41. Shim, M., Wilson, B., Marple, E., and Wach, M., Study of fiber-optic probes for in vivo medicalRaman spectroscopy, Appl. Spectrosc., 53, 619, 1999.

42. Motz, J.T., Hunter, M., Galindo, L., Kramer, J.R., Dasari, R.R., and Feld, M.S., Development ofoptical fiber probes for biological Raman spectroscopy, in Biomedical Optical Spectroscopy andDiagnostics, OSA Biomedical Topical Meetings Technical Digest, 336, Optical Society of America,Washington, D.C., 2002.

43. Yu, N.-T., DeNagel, D.C., Pruett, P.L., and Kuck, J.F.R., Jr., Disulfide bond formation in the eyelens, Proc. Natl. Acad. Sci. U.S.A., 82, 7965, 1985.

44. Palmer, G.M., Marshek, C.L., Vrotsos, K.M., and Ramanujam, N., Optimal methods for fluores-cence and diffuse reflectance measurements of tissue biopsy samples, Lasers Surg. Med., 30, 191,2002.

45. Williams, A.C., Barry, B.W., Edwards, H.G., and Farwell, D.W., A critical comparison of someRaman spectroscopic techniques for studies of human stratum corneum, Pharm. Res., 10, 1642,1993.

46. ANSI, American National Standards for Safe Use of Lasers, American National Safety Institute,Orlando, FL, 2000.

47. Lakowicz, J.R., Principles of Fluorescence Spectroscopy, Plenum Press, New York, 1983.48. Van Duyne, R.P., Jeanmaire, D.L., and Shriver, D.F., Mode-locked laser Raman spectroscopy: a new

technique for the rejection of interfering background luminescence signals. Anal. Chem., 46, 213,1974.

49. Mosier-Boss, P.A., Lieberman, S.H., and Newberry, R., Fluorescence rejection in Raman spectros-copy by shifted-spectra, edge detection, and FFT filtering techniques, Appl. Spectrosc., 49, 630, 1995.

50. Zhang, D.M. and Ben-Amotz, D., Enhanced chemical classification of Raman images in the pres-ence of strong fluorescence interference, Appl. Spectrosc., 54, 1379, 2000.

51. O’Grady, A., Dennis, A.C., Denvir, D., McGarvey, J.J., and Bell, S.E.J., Quantitative Raman spec-troscopy of highly fluorescent samples using pseudosecond derivatives and multivariate analysis,Anal. Chem., 73, 2058, 2001.

52. Barclay, V.J., Bonner, R.F., and Hamilton, I.P., Application of wavelet transforms to experimentalspectra: smoothing, denoising, and data set compression, Anal. Chem., 69, 78, 1997.

53. Cai, T.T., Zhang, D.M., and Ben-Amotz, D., Enhanced chemical classification of Raman imagesusing multiresolution wavelet transformation, Appl. Spectrosc., 55, 1124, 2001.

54. Vickers, T.J., Wambles, R.E., and Mann, C.K., Curve fitting and linearity: data processing in Ramanspectroscopy, Appl. Spectrosc., 55, 389, 2001.

55. Baraga, J.J., Feld, M.S., and Rava, R.P., Rapid near-infrared Raman spectroscopy of human tissuewith a spectrograph and CCD detector, Appl. Spectrosc., 46, 187, 1992.

56. Lieber, C.A., Molpus, K., Brader, K., and Mahadevan-Jansen, A., Diagnostic tool for early detectionof ovarian cancers using Raman spectroscopy, in Biomedical Spectroscopy: Vibrational Spectroscopyand Other Novel Techniques, Mahadevan-Jansen, A. and Puppels, G.J., Eds., 3918, SPIE, Bellingham,WA, 2000.

57. Liu, C.H., Das, B.B., Sha Glassman, W.L., Tang, G.C., Yoo, K.M., Zhu, H.R., Akins, D.L., Lubicz,S.S., Cleary, J., Prudente, R., Cellmer, E., Caron, A., and Alfano, R.R., Raman, fluorescence andtime-resolved light scattering as optical diagnostic techniques to separate diseased and normalbiomedical media, J. Photochem. Photobiol. B, 16, 187, 1992.

©2003 CRC Press LLC

Page 26: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

58. Mahadevan-Jansen, A., Mitchell, M.F., Ramanujam, N., Malpica, A., Thomsen, S., Utzinger, U.,and Richards-Kortum, R., Near-infrared Raman spectroscopy for in vitro detection of cervicalprecancers, Photochem. Photobiol., 68, 123, 1998.

59. Berger, A.J., Koo, T.W., Itzkan, I., Horowitz, G., and Feld, M.S., Multicomponent blood analysisby near-infrared Raman spectroscopy, Appl. Opt., 38, 2916, 1999.

60. Shafer-Peltier, K.E., Haka, A.S., Fitzmaurice, M., Crowe, J., Myles, J., Dasari, R.R., and Feld, M.S.,Raman microspectroscopic model of human breast tissue: implications for breast cancer diagnosisin vivo, J. Raman Spectrosc., 33, 552, 2002.

61. Koljenovic, S., Choo-Smith, L.P., Bakker Schut, T.C., Kros, J.M., van den Berge, H.J., and Puppels,G.J., Discriminating vital tumor from necrotic tissue in human glioblastoma tissue samples byRaman spectroscopy, Lab Invest., 82, 1265, 2002.

62. Caspers, P.J., Lucassen, G.W., Carter, E.A., Bruining, H.A., and Puppels, G.J., In vivo confocalRaman microspectroscopy of the skin: noninvasive determination of molecular concentrationprofiles, J. Invest. Dermatol., 116, 434, 2001.

63. Stone, N., Kendall, C., Shepherd, N., Crow, P., and Barr, H., Near-infrared Raman spectroscopyfor the classification of epithelial pre-cancers and cancers, J. Raman Spectrosc., 33, 564, 2002.

64. Nijssen, A., Bakker Schut, T.C., Heule, F., Caspers, P.J., Hayes, D.P., Neumann, M.H., and Puppels,G.J., Discriminating basal cell carcinoma from its surrounding tissue by Raman spectroscopy,J. Invest. Dermatol., 119, 64, 2002.

65. Cancer Facts and Figures 2002, American Cancer Society, New York, 2002.66. Pearce, E., Anatomy and Physiology for Nurses, Oxford University Press, New York, 1989.67. Robbins, S.L., Cotran, R.S., and Kumar, V., Pathologic Basis of Disease, W.B. Saunders, Philadelphia,

1994.68. Manoharan, R., Shafer, K., Perelman, L., Wu, J., Chen, K., Deinum, G., Fitzmaurice, M., Myles, J.,

Crowe, J., Dasari, R.R., and Feld, M.S., Raman spectroscopy and fluorescence photon migrationfor breast cancer diagnosis and imaging, Photochem. Photobiol., 67, 15, 1998.

69. Centeno, J.A., Mullick, F.G., Panos, R.G., Miller, F.W., and Valenzuela-Espinoza, A., Laser-Ramanmicroprobe identification of inclusions in capsules associated with silicone gel breast implants,Mod. Pathol., 12, 714, 1999.

70. Ferenczy, A. and Winkler, B., Cervical intraepithelial neoplasia, in Pathology of the Female GenitalTract, Blaustein, A., Ed., Springer-Verlag, New York, 1982, p. 156.

71. Bosch, F.X., Munoz, N., and de Sanjose, S., Human papillomavirus and other risk factors forcervical cancer, Biomed. Pharmacother., 51, 268, 1997.

72. Myers, E.R., McCrory, D.C., Subramanian, S., McCall, N., Nanda, K., Datta, S., and Matchar, D.B.,Setting the target for a better cervical screening test: characteristics of a cost-effective test for cervicalneoplasia screening, Obstet. Gynecol., 96, 645, 2000.

73. Robichaux, A., Shappell, H., Huff, B., Jones, H., and Mahadevan-Jansen, A., In vivo detection ofcervical dysplasia using near infrared Raman spectroscopy, Lasers Surg. Med., Suppl. 14, 5, 2002.

74. Sterenborg, N.J., Thomsen, S., Jacques, S.L., Duvic, M., Motamedi, M., and Wagner, R.F., Jr., Invivo fluorescence spectroscopy and imaging of human skin tumors, Dermatol. Surg., 21, 821, 1995.

75. Panjehpour, M., Julius, C.E., Phan, M.N., Vo-Dinh, T., and Overholt, S., Laser-induced fluorescencespectroscopy for in vivo diagnosis of non-melanoma skin cancers, Lasers Surg. Med., 31, 367, 2002.

76. Williams, A.C., Edwards, H.G.M., and Barry, B.W., Fourier-transform Raman spectroscopy — anovel application for examining human stratum corneum, Intl. J. Pharm., 81, R11, 1992.

77. Williams, A.C., Edwards, H.G.M., and Barry, B.W., Raman spectra of human keratotic biopolymers:skin, callus, hair and nail, J. Raman Spectrosc., 25, 95, 1994.

78. Hata, T.R., Scholz, T.A., Ermakov, I.V., McClane, R.W., Khachik, F., Gellermann, W., and Pershing,L.K., Non-invasive Raman spectroscopic detection of carotenoids in human skin, J. Invest. Derma-tol., 115, 441, 2000.

©2003 CRC Press LLC

Page 27: Chapter 30: Raman Spectroscopy From Benchtop to …huynhqlinh/qpys/seminar/Seminar2005/...in the near-infrared region of the spectrum where fluorescence is reduced, have resulted

79. Gniadecka, M., Wulf, H.C., Nielsen, O.F., Christensen, D.H., and Hercogova, J., Distinctive molec-ular abnormalities in benign and malignant skin lesions: studies by Raman spectroscopy, Photo-chem. Photobiol., 66, 418, 1997.

80. Lieber, C., Robichaux, A., Ellis, D., and Mahadevan-Jansen, A., Detection of skin abnormalitiesusing near infrared confocal Raman spectroscopy, Lasers Surg. Med., Suppl. 14, 4, 2002.

81. Wolthuis, R., van Aken, M., Fountas, K., Robinson, J.S., Jr., Bruining, H.A., and Puppels, G.J.,Determination of water concentration in brain tissue by Raman spectroscopy, Anal. Chem., 73,3915, 2001.

82. Boustany, N.N., Manoharan, R., and Dasari, R.R., Ultraviolet resonance Raman spectroscopy ofbulk and microscopic human colon tissue, Appl. Spectrosc., 54, 24, 2000.

83. Peterson, C.M., Peterson, K.P., Chaiken, J., Finney, W., Knudson, P.E., and Weinstock, R.S., Non-invasive monitoring of blood glucose using Raman spectroscopy, Diabetes, 49, 490, 2000.

©2003 CRC Press LLC