raman spectroscopy substrate optimization and method

157
University of Tennessee, Knoxville University of Tennessee, Knoxville TRACE: Tennessee Research and Creative TRACE: Tennessee Research and Creative Exchange Exchange Doctoral Dissertations Graduate School 5-2020 Raman Spectroscopy Substrate Optimization and Method Raman Spectroscopy Substrate Optimization and Method Development for the Detection and Analysis of Cyclic Organic Development for the Detection and Analysis of Cyclic Organic Molecules and Materials Molecules and Materials Terence J. Riffey-Moore University of Tennessee Follow this and additional works at: https://trace.tennessee.edu/utk_graddiss Recommended Citation Recommended Citation Riffey-Moore, Terence J., "Raman Spectroscopy Substrate Optimization and Method Development for the Detection and Analysis of Cyclic Organic Molecules and Materials. " PhD diss., University of Tennessee, 2020. https://trace.tennessee.edu/utk_graddiss/5878 This Dissertation is brought to you for free and open access by the Graduate School at TRACE: Tennessee Research and Creative Exchange. It has been accepted for inclusion in Doctoral Dissertations by an authorized administrator of TRACE: Tennessee Research and Creative Exchange. For more information, please contact [email protected].

Upload: others

Post on 15-Feb-2022

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Raman Spectroscopy Substrate Optimization and Method

University of Tennessee, Knoxville University of Tennessee, Knoxville

TRACE: Tennessee Research and Creative TRACE: Tennessee Research and Creative

Exchange Exchange

Doctoral Dissertations Graduate School

5-2020

Raman Spectroscopy Substrate Optimization and Method Raman Spectroscopy Substrate Optimization and Method

Development for the Detection and Analysis of Cyclic Organic Development for the Detection and Analysis of Cyclic Organic

Molecules and Materials Molecules and Materials

Terence J. Riffey-Moore University of Tennessee

Follow this and additional works at: https://trace.tennessee.edu/utk_graddiss

Recommended Citation Recommended Citation Riffey-Moore, Terence J., "Raman Spectroscopy Substrate Optimization and Method Development for the Detection and Analysis of Cyclic Organic Molecules and Materials. " PhD diss., University of Tennessee, 2020. https://trace.tennessee.edu/utk_graddiss/5878

This Dissertation is brought to you for free and open access by the Graduate School at TRACE: Tennessee Research and Creative Exchange. It has been accepted for inclusion in Doctoral Dissertations by an authorized administrator of TRACE: Tennessee Research and Creative Exchange. For more information, please contact [email protected].

Page 2: Raman Spectroscopy Substrate Optimization and Method

To the Graduate Council:

I am submitting herewith a dissertation written by Terence J. Riffey-Moore entitled "Raman

Spectroscopy Substrate Optimization and Method Development for the Detection and Analysis

of Cyclic Organic Molecules and Materials." I have examined the final electronic copy of this

dissertation for form and content and recommend that it be accepted in partial fulfillment of the

requirements for the degree of Doctor of Philosophy, with a major in Chemistry.

Bhavya Sharma, Major Professor

We have read this dissertation and recommend its acceptance:

Zi-Ling Xue, Sharani Roy, Luca Giori

Accepted for the Council:

Dixie L. Thompson

Vice Provost and Dean of the Graduate School

(Original signatures are on file with official student records.)

Page 3: Raman Spectroscopy Substrate Optimization and Method

Raman Spectroscopy Substrate Optimization and Method Development

for the Detection and Analysis of Cyclic Organic Molecules and

Materials

A Dissertation Presented for the

Doctor of Philosophy

Degree

The University of Tennessee, Knoxville

Terence Joshua Riffey-Moore

May 2020

Page 4: Raman Spectroscopy Substrate Optimization and Method

ii

Copyright © 2020 by Terence Joshua Riffey-Moore All rights reserved.

Page 5: Raman Spectroscopy Substrate Optimization and Method

iii

DEDICATION

To my mother, Diana, who from my earliest memories taught me the power of an

education and where it would take me. I couldn’t be more grateful for the love and

support you’ve given me during this time, and my whole life. I hope I’ve made you

proud.

To my husband, Beau. I absolutely could not have done any of this without you. Thank

you for being my champion, cheerleader, and sounding board when I couldn’t figure

something out. I can’t wait to experience our lives after all of this is in the rear-view

mirror, and we’re setting out on our own adventure with Isaiah. I love you tremendously

and thank you for your love and support.

Page 6: Raman Spectroscopy Substrate Optimization and Method

iv

ACKNOWLEDGEMENTS

First and foremost, I would like to thank God, through whom all things are possible.

Thank you to my advisor, Dr. Bhavya Sharma, for making me a better scientist than I could

have imagined going into this process. You have always been there to provide me the

necessary motivation, and I thank you for everything you have done for me these past 5

years.

To my current and former group members, you all have helped make even my worst

and most panicked days bearable with your compassion, comfort, and humor. I’ve been

asked what I’ll miss the most about this experience, and the answer always comes back to

that I will miss working with each of you every day.

To the faculty, staff, and other graduate students of the Chemistry department,

thank you for being my surrogate family for the last 5 years. I look forward to seeing all

the amazing things you’ll accomplish down the road!

To my family, I thank you for your patience with me while I’ve gone on this

adventure. I have always been motivated to pursue my educational dreams because I was

taught at a very early age that education is the most important thing you’ll ever acquire.

Especially to my mother, Diana, I am thankful for everything you have done and continue

to do for me. I hope I make you as proud that I am your son as I am to call you my mother.

To my husband, Beau, I want to acknowledge your patience with this process, and

thank you for making me want to be a better person each morning when I wake up. I’m so

blessed to get to go through this life with you, and I can’t wait to see the doors open for us

in the future.

Page 7: Raman Spectroscopy Substrate Optimization and Method

v

ABSTRACT

Organic molecules are ubiquitous species, with approximately 9 million

compounds classified as organic molecules, and more being synthesized in laboratories

around the world each day. Of these molecules, cyclic organic molecules are especially

important in biological and industrial settings. Detection of these molecules is of

paramount interest in several fields, including biosensing, environmental, and process

monitoring. Additionally, materials are made of cyclic organic molecules, and they can be

characterized using a variety of techniques. Raman spectroscopy is an attractive method

for both the detection of molecules and characterization of materials.

Raman scattering is a phenomenon in which light is inelastically scattered by a

molecule upon interaction with monochromatic light. The inelastically scattered photons

differ in energy from the incident photons by the magnitude of a molecular vibrational

mode frequency. The resulting spectrum is considered a molecular “fingerprint,” since each

molecule gives a unique Raman spectrum based on the constituent atoms and bonding

environments. This excellent specificity allows Raman scattering to be used for a variety

of detection and characterization processes.

Raman scattering is an inherently weak phenomenon, but this weakness can be

overcome by placing the molecule of interest within 1 – 2 nm of a plasmonic metal surface.

The incident light causes the conduction band electrons of the metal to oscillate, which

results in an increased electric field at the surface due to the localized surface plasmon

resonance (LSPR) effect, which then increases the intensity of the Raman scattered light.

The use of these plasmonic metal surfaces for amplified Raman scattering is called surface

Page 8: Raman Spectroscopy Substrate Optimization and Method

vi

enhanced Raman spectroscopy (SERS). This surface enhancement allows for trace levels

of analyte molecules to be detected.

Experiments herein were conducted to detect cortisol at physiological

concentrations using silver colloidal nanoparticles. A detection limit of 177 nM was

established, which falls at the low end of the physiological concentration. Additionally,

cyclic volatile organic compounds were detected in the gas phase using a multi-

dimensional SERS substrate, which shows detection of benzenethiol at the parts-per-

million level. Lastly, normal Raman spectroscopy is used to determine the surface

characteristics and interior character of a series of oxidized polyacrylonitrile-based carbon

fibers.

Page 9: Raman Spectroscopy Substrate Optimization and Method

vii

TABLE OF CONTENTS

Chapter 1 Introduction..................................................................................................... 1 1.1 Introduction ............................................................................................................... 2 1.2 References ................................................................................................................. 5

Chapter 2 Raman Theory and Instrumentation .......................................................... 10 2.1 Raman Spectroscopy ............................................................................................... 11 2.2 Raman Theory ......................................................................................................... 13 2.3 Surface-Enhanced Raman Spectroscopy ................................................................ 18 2.4 Enhancing Substrate Synthesis and Fabrication ..................................................... 26

2.4.1 Gold Nanoparticle Synthesis ............................................................................ 26 2.4.2 Silver Nanoparticle Synthesis .......................................................................... 27 2.4.3 2D/3D SERS Substrates ................................................................................... 28

2.5 Enhancement Factors .............................................................................................. 30 2.6 Raman Instrumentation ........................................................................................... 31

2.6.1 Excitation Sources ........................................................................................... 31 2.6.2 Samples and Collection Optics ........................................................................ 33 2.6.3 Dispersion System ........................................................................................... 34 2.6.4 Detectors .......................................................................................................... 35 2.6.5 Raman Instrumentation in the Sharma Lab ..................................................... 36

2.7 References ............................................................................................................... 40 Chapter 3 Direct Surface Enhanced Raman Spectroscopic Detection of Cortisol at Physiological Concentrations ......................................................................................... 45

3.1 Abstract ................................................................................................................... 47 3.2 Introduction ............................................................................................................. 48 3.3 Experimental Section .............................................................................................. 50

3.3.1 Materials and Reagents .................................................................................... 50 3.3.2 Synthesis of Hydroxylamine-reduced Silver Nanoparticles ............................ 50 3.3.3 SERS Sample Preparation................................................................................ 51 3.3.4 Normal Raman and SERS Spectral Acquisition .............................................. 54 3.3.5 Theoretical Raman Calculations ...................................................................... 54

3.4 Results and Discussion ........................................................................................... 55 3.5 Conclusions ............................................................................................................. 72 3.6 Acknowledgements ................................................................................................. 72 3.7 References ............................................................................................................... 73

Chapter 4 Surface Enhanced Raman-based Detection of Volatile Organic Compounds Using Porous Silicon Oxide Coated Disc-on-Pillar Arrays ................... 79

4.1 Abstract ................................................................................................................... 81 4.2 Introduction ............................................................................................................. 81 4.3 Experimental Methods ............................................................................................ 81

4.3.1 Reagents ........................................................................................................... 84 4.3.2 Nanofabrication................................................................................................ 84 4.3.3 Off-gassing Experiments and SERS Spectra Collection ................................. 85

4.4 Results and Discussion ........................................................................................... 85 4.5 Conclusions ........................................................................................................... 100

Page 10: Raman Spectroscopy Substrate Optimization and Method

viii

4.6 Acknowledgements ............................................................................................... 100 4.7 References ............................................................................................................. 101

Chapter 5 Normal Raman Analysis of the Composition of Oxidized Polyacrylonitrile Carbon Fibers................................................................................................................ 108

5.1 Abstract ................................................................................................................. 109 5.2 Introduction ........................................................................................................... 111 5.3 Methods................................................................................................................. 112

5.3.1 Samples .......................................................................................................... 113 5.3.2 Raman Instrumentation .................................................................................. 113 5.3.3 Spectral Collection ......................................................................................... 115 5.3.4 Spectral Processing ........................................................................................ 115

5.4 Results and Discussion ......................................................................................... 116 5.5 Conclusions ........................................................................................................... 134 5.6 Acknowledgements ............................................................................................... 136 5.7 References ............................................................................................................. 137

Chapter 6 Conclusions and Future Directions ........................................................... 140 VITA............................................................................................................................... 143

Page 11: Raman Spectroscopy Substrate Optimization and Method

ix

LIST OF FIGURES

Figure 2.1. Jablonski diagram for Rayleigh and Raman scattering. E0 is the electronic ground state energy, E1 is the first excited electronic state energy, h is Planck’s constant, and ν0 and νm are the frequencies of the ground vibrational state and a vibrational mode, respectively. ................................................................................. 12

Figure 2.2. A typical Raman spectrum obtained from neat pyridine, a strong Raman scatterer. The spectrum is plotted as a with intensity as a function of Raman shift, and was obtained in 1 s using 633 nm excitation at 5.7 mW of power. ................... 16

Figure 2.3. The localized surface plasmon resonance (LSPR) phenomenon, in which conduction band electrons of a metal nanostructure oscillate in the presence of an external electric field................................................................................................. 20

Figure 2.4. Theoretical simulations of the EM field enhancement around silver nanoparticles, with single triangular (top left) and ellipsoidal morphologies (right), as well as dimers of spherical nanoparticles (bottom left), shown. .......................... 21

Figure 2.5. Approximate wavelength ranges where Ag, Au, and Cu have been well-characterized and are established to support SERS. ................................................. 24

Figure 2.6. SEM image (a) and extinction spectrum (b) of colloidal gold synthesized by the Frens method. ...................................................................................................... 27

Figure 2.7. Schematic of the home-built Raman microscope system in the Sharma Lab, showing 4 excitation source options and indicating the beam path for each source. 39

Figure 3.1. SEM image of the prepared hydroxylamine-reduced silver nanoparticles (HA-AgNP), showing good monodispersity and an average diameter of ~ 30 nm. . 52

Figure 3.2. Extinction spectrum of synthesized HA-AgNP, showing an LSPR maximum located at 409 nm. ..................................................................................................... 53

Figure 3.3. Structure of cortisol (a) and the carbon-numbering and ring identification conventions for the cholesterol backbone present in cortisol (b). ............................ 56

Figure 3.4. Theoretical (a) and experimental (b) normal Raman spectra of cortisol solid in the region from 200−1800 cm−1. There is good agreement between the theoretical and experimental spectra. For the experimental spectrum, λex = 532 nm, P = 10 mW, and t = 10 s. .............................................................................................. 57

Figure 3.5. Experimental (top, red) and theoretical (bottom, blue) normal Raman spectra for cortisol in the region from 200 – 4000 cm-1. For the experimental spectrum, λex = 532 nm, P = 10 mW, and t = 10 s. ......................................................................... 58

Figure 3.6. Theoretical (a) and experimental (b) normal Raman spectra for cortisol in the high wavenumber region........................................................................................... 59

Figure 3.7. SERS spectra (1000 – 1600 cm-1) of nM cortisol solutions representative of the physiological normal range in serum. Spectra were collected with λex = 532 nm, P = 10 mW, t = 30 s. ................................................................................................. 62

Figure 3.8. Structure of methylene blue. .......................................................................... 64 Figure 3.9. Sigmoidal concentration response plots obtained after peak fitting triplicate

SERS spectra of physiological cortisol concentrations for the 1269 cm-1 (top) and 1504 cm-1 (bottom) peaks. ........................................................................................ 66

Figure 3.10. Sigmoidal concentration response plots for the observed peaks in SERS spectra of cortisol obtained from samples in the physiological range (125-1000 nM).

Page 12: Raman Spectroscopy Substrate Optimization and Method

x

Error bars are present on all data points, but the small magnitude of some error bars is obscured by the data marker in several instances. ................................................ 68

Figure 3.11. Linear regression plots for the observed peaks in SERS spectra of cortisol obtained from samples in the physiological range (125-1000 nM). ......................... 69

Figure 3.12. Principal components analysis of cortisol in the physiologically relevant concentration range (low μM to 125 nM) in 0.75% bovine serum albumin-phosphate buffered saline (BSA-PBS). (a) Plot of scores for PC1 vs. PC2. The data point labels indicate the concentration of cortisol present in the sample. (b) The loading for PC1, which shows that PC1 is dominated by contributions from the BSA. (c) The loading for PC2, which shows peaks arising from cortisol. .................................................. 71

Figure 4.1. SERS spectra obtained from substrates (with and without pillar architecture) that were exposed to the off-gassed vapor from neat benzenethiol for 10 minutes. λex = 785 nm, P = 5 mW, t = 60 s ................................................................................... 87

Figure 4.2. SERS activity for pre- and post-alumina layer deposition substrates. Spectral analysis shows that a 40% decrease in signal, but the signal obtained after deposition is still reasonable for analysis. λex = 785 nm, P = 5 mW, t = 30 s ............................ 89

Figure 4.3. SERS spectra obtained from 10-minute exposure time to off-gassed neat benzenethiol vapor, demonstrating the best enhancement for the substrate occurs with 1000 nm of porous silicon oxide (PSO). λex = 785 nm, P = 5 mW, t = 60 s .... 91

Figure 4.4. SERS spectra obtained after off-gassing a) neat benzenethiol and b) neat pyridine onto optimized Au disc-on-pillar arrays with 1000 nm of porous silicon oxide demonstrating that it is possible to rapidly detect either analyte with as little as 30 s of exposure to the vapor. λex = 785 nm, P = 5 mW, t = 1 s ............................... 92

Figure 4.5. SERS spectra obtained after 10 minutes of off gassing from a) 100 ppm, b) 10 ppm, and c) 1 ppm benzenethiol in water. λex = 785 nm, P = 5 mW, t = 10 s ..... 94

Figure 4.6. Linear regression analysis of integrated intensity of the 1572 cm-1 peak obtained from off gassing aqueous parts-per-million solutions of benzenethiol. ..... 95

Figure 4.7. Linear regression analysis of the (a) 996, (b) 1020, (c) 1070, and (d) 1572 cm-1 peaks obtained from the off-gassing of parts-per-million aqueous solutions of benzenethiol. ............................................................................................................. 96

Figure 4.8. SERS spectra obtained after Au-DOP substrate was kept in contact with vapors off-gassed from pools of a) neat benzene and b) ortho-xylene. λex = 785 nm, P = 5 mW, t = 60 s .................................................................................................... 98

Figure 5.1. Schematic of the Raman apparatus used for the analysis of the oxidized PAN fibers. ...................................................................................................................... 114

Figure 5.2. The first-order Raman spectra (1000 – 1800 cm-1) obtained from fibers 1 – 11, showing differences in the spectral lineshapes and line widths for the fiber samples. ................................................................................................................... 117

Figure 5.3. The Raman spectrum obtained from fiber 1 (a) and the fitting of the D- and G-band of the fiber (b). The D-band is fitted with a Lorentzian lineshape, whereas the G-band is fitted using a predominantly Gaussian lineshape. ............................ 118

Figure 5.4. Results of fiber 11 cross-section Raman analysis: (a, d) Raman spectra collected at 1 μm intervals across cross sections of fiber 11. (b, e) The integrated intensity of the D- and G-bands at each collection point. (c, f) The D/G ratio at each collection point........................................................................................................ 121

Page 13: Raman Spectroscopy Substrate Optimization and Method

xi

Figure 5.5. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 1. ..................................................................................................................... 123

Figure 5.6. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 2. ..................................................................................................................... 125

Figure 5.7. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 3. ..................................................................................................................... 126

Figure 5.8. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 4. ..................................................................................................................... 127

Figure 5.9. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 5. ..................................................................................................................... 128

Figure 5.10. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 6. ..................................................................................................................... 129

Figure 5.11. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 7. ..................................................................................................................... 131

Figure 5.12. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 8. ..................................................................................................................... 132

Figure 5.13. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 9. ..................................................................................................................... 133

Figure 5.14. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 10. ................................................................................................................... 135

Page 14: Raman Spectroscopy Substrate Optimization and Method

1

Chapter 1

Introduction

Page 15: Raman Spectroscopy Substrate Optimization and Method

2

1.1 Introduction

Raman spectroscopy (RS) is a widely used vibrational spectroscopy technique

which offers excellent chemical specificity, with the resulting spectra being unique to the

molecule of interest. Because of this high level of chemical specificity, RS has been

employed to extensively probe systems to extract chemical information from them.

However, Raman scattering is weak compared to other scattering processes, meaning that

signal must be amplified for the purpose of trace analyte detection. A common way to

overcome this deficiency is through the use of surface-enhanced RS (SERS), which

advantageously relies on the generation of enhanced electric fields at the surface of a

nanostructure due to the localized surface plasmon resonance (LSPR) effect, resulting in

increased Raman scattering from samples adsorbed onto or close to the surface.

In biological matrices, analytes are usually present at ultralow concentrations,

which makes employing SERS an attractive technique for study of biological systems.

There are several review papers that are focused on the use of RS techniques in biological

and biomedical applications.1-5 The increased sensitivity provided by SERS coupled with

the inherent chemical specificity of normal RS allows for the detection of these biologically

relevant analytes, including biomarkers of a wide variety of disease states,6 glucose,7 and

neurotransmitters.8 Neurotransmitters detection has also been demonstrated through the

skull by Moody et al. using surface-enhanced spatially-offset RS (SESORS).9, 10 However,

RS is not limited to small molecule bioanalysis, as it has been used to characterize larger

biomolecules, including lipids11, 12 and proteins,13-16 all the way up to complete cells.17-19

Page 16: Raman Spectroscopy Substrate Optimization and Method

3

RS variants have also demonstrated extreme versatility in the field of environmental

monitoring. Gajaraj et al. have employed SERS based methods to determine nitrate

concentrations in wastewater at 0.5 mg L-1 limits of detection while simultaneously

detecting sulfate,20 and these compounds have also been detected by Ianoul et al. at <200

parts-per-billion using deep-UV resonance Raman spectroscopy methods.21 Nitrates have

detrimental effects on aquatic ecosystems, including the heightened eutrophication of both

freshwater and marine environments and methemoglobinemia in both aquatic species and

humans.22 Microplastics, which are pieces of various plastics (e.g. polyethylene, polyvinyl

chloride, or polystyrene) that measure smaller than 5 mm, are an increasingly prevalent

environmental pollutant, with studies on the amount and identity of these microplastics

carried out in Asia, Europe, Oceania, Africa, and North America.23 RS has been employed

as a method to confirm visual identification of microplastic identity in water samples24 and

stomachs of plankton-ingesting fishes.25 Another toxic pollutant, 1,2,3-trichloropropane,

has been detected in water at sub-millimolar concentrations using an optofluidic SERS

substrate by Pilát et al.26 The technique has also been used in detection of azo dyes, which

comprise about 50% of the dye molecules produced worldwide and have been shown to

have carcinogenic properties, and nuclear waste testing.14, 27 As demonstrated, a wide

variety of these environmental pollutants can be detected using RS and variants.

Aside from outright detection of an analyte molecule of interest, RS can be used to

characterize the microstructure of a material when coupled with a microscope that can

reduce the spot size to the order of micrometers. This materials characterization approach

has been used extensively to characterize carbonaceous materials, but additional

applications include the characterization of triangular silicon nanowires via normal RS,28

Page 17: Raman Spectroscopy Substrate Optimization and Method

4

and the analysis of functional groups in coal samples.29 Plasmon-enhanced RS techniques

have been used to analyze the growth kinetics of liquid-phase deposition of CdS thin

films,30 study the ion-intercalating processes in electrodes,31 and to evaluate ferroelectric

domains in barium titanate (BaTiO3) nanorods.32 These techniques can also be applied to

biological materials, with Heufner et al. demonstrating cellular mapping of endocytosis

pathways.33

Contained herein is an example of the application of SERS to detect a biologically

relevant molecule (Chapter 3) and to detect gaseous analytes that are byproducts of energy

reactions and pose substantial threat to the environment and human health (Chapter 4).

Lastly, Chapter 5 deals with the application of RS as a method to characterize and analyze

the microstructure of carbon fiber materials. In each of these chapters, the sensing substrate

was optimized, and the method was tailored to the specific analyte molecule.

Page 18: Raman Spectroscopy Substrate Optimization and Method

5

1.2 References

1. Ember, K. J. I.; Hoeve, M. A.; McAughtrie, S. L.; Bergholt, M. S.; Dwyer, B.

J.; Stevens, M. M.; Faulds, K.; Forbes, S. J.; Campbell, C. J., Raman spectroscopy and

regenerative medicine: a review. NPJ Regen. Med. 2017, 2 (1), 12.

2. Laing, S.; Jamieson, L. E.; Faulds, K.; Graham, D., Surface-enhanced Raman

spectroscopy for in vivo biosensing. Nat. Rev. Chem. 2017, 1 (8).

3. Lorenz, B.; Wichmann, C.; Stöckel, S.; Rösch, P.; Popp, J., Cultivation-Free

Raman Spectroscopic Investigations of Bacteria. Trends Microbiol. 2017, 25 (5), 413-

424.

4. Shen, Y.; Hu, F.; Min, W., Raman Imaging of Small Biomolecules. Annu. Rev.

Biophys. 2019, 48 (1), 347-369.

5. Smith, R.; Wright, K. L.; Ashton, L., Raman spectroscopy: an evolving technique

for live cell studies. Analyst 2016, 141 (12), 3590-600.

6. Moore, T.; Moody, A.; Payne, T.; Sarabia, G.; Daniel, A.; Sharma, B., In Vitro

and In Vivo SERS Biosensing for Disease Diagnosis. Biosensors 2018, 8 (2).

7. Sharma, B.; Bugga, P.; Madison, L. R.; Henry, A.-I.; Blaber, M. G.;

Greeneltch, N. G.; Chiang, N.; Mrksich, M.; Schatz, G. C.; Van Duyne, R. P.,

Bisboronic Acids for Selective, Physiologically Relevant Direct Glucose Sensing with

Surface-Enhanced Raman Spectroscopy. J. Am. Chem. Soc. 2016, 138 (42), 13952-

13959.

Page 19: Raman Spectroscopy Substrate Optimization and Method

6

8. Moody, A. S.; Sharma, B., Multi-metal, Multi-wavelength Surface-Enhanced

Raman Spectroscopy Detection of Neurotransmitters. ACS Chem. Neurosci. 2018, 9 (6),

1380-1387.

9. Moody, A. S.; Baghernejad, P. C.; Webb, K. R.; Sharma, B., Surface Enhanced

Spatially Offset Raman Spectroscopy Detection of Neurochemicals Through the Skull.

Anal. Chem. 2017, 89 (11), 5688-5692.

10. Moody, A. S.; Payne, T. D.; Barth, B. A.; Sharma, B., Surface-enhanced

spatially-offset Raman spectroscopy (SESORS) for detection of neurochemicals through

the skull at physiologically relevant concentrations. Analyst 2020, 145 (5), 1885-1893.

11. Czamara, K.; Majzner, K.; Pacia, M. Z.; Kochan, K.; Kaczor, A.; Baranska, M.,

Raman spectroscopy of lipids: a review. J. Raman Spectrosc. 2015, 46 (1), 4-20.

12. Krafft, C.; Neudert, L.; Simat, T.; Salzer, R., Near infrared Raman spectra of

human brain lipids. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2005, 61 (7), 1529-35.

13. Benevides, J. M.; Overman, S. A.; Thomas, G. J., Jr., Raman spectroscopy of

proteins. Curr. Protoc. Protein Sci. 2004, Chapter 17 (1), Unit 17 8.

14. Das, R. S.; Agrawal, Y. K., Raman spectroscopy: Recent advancements,

techniques and applications. Vib. Spectrosc. 2011, 57 (2), 163-176.

15. Rygula, A.; Majzner, K.; Marzec, K. M.; Kaczor, A.; Pilarczyk, M.; Baranska,

M., Raman spectroscopy of proteins: a review. J. Raman Spectrosc. 2013, 44 (8), 1061-

1076.

16. Sharma, B.; Bykov, S. V.; Asher, S. A., UV Resonance Raman Investigation of

Electronic Transitions in α-Helical and Polyproline II-Like Conformations. J. Phys.

Chem. B 2008, 112 (37), 11762-11769.

Page 20: Raman Spectroscopy Substrate Optimization and Method

7

17. Chan, J. W.; Lieu, D. K.; Huser, T.; Li, R. A., Label-free separation of human

embryonic stem cells and their cardiac derivatives using Raman spectroscopy. Anal.

Chem. 2009, 81 (4), 1324-31.

18. Chen, F.; Flaherty, B. R.; Cohen, C. E.; Peterson, D. S.; Zhao, Y., Direct

detection of malaria infected red blood cells by surface enhanced Raman spectroscopy.

Nanomedicine 2016, 12 (6), 1445-51.

19. Premasiri, W. R.; Lee, J. C.; Sauer-Budge, A.; Théberge, R.; Costello, C. E.;

Ziegler, L. D., The biochemical origins of the surface-enhanced Raman spectra of

bacteria: a metabolomics profiling by SERS. Anal. Bioanal. Chem. 2016, 408 (17), 4631-

4647.

20. Gajaraj, S.; Fan, C.; Lin, M.; Hu, Z., Quantitative detection of nitrate in water

and wastewater by surface-enhanced Raman spectroscopy. Environ. Monit. Assess. 2012,

185 (7), 5673-5681.

21. Ianoul, A.; Coleman, T.; Asher, S. A., UV Resonance Raman Spectroscopic

Detection of Nitrate and Nitrite in Wastewater Treatment Processes. Anal. Chem. 2002,

74 (6), 1458-1461.

22. Gajaraj, S.; Fan, C.; Lin, M.; Hu, Z., Quantitative detection of nitrate in water

and wastewater by surface-enhanced Raman spectroscopy. Environ. Monit. Assess. 2013,

185 (7), 5673-81.

23. Qiu, Q.; Tan, Z.; Wang, J.; Peng, J.; Li, M.; Zhan, Z., Extraction, enumeration

and identification methods for monitoring microplastics in the environment. Estuar.

Coast. Shelf Sci. 2016, 176, 102-109.

Page 21: Raman Spectroscopy Substrate Optimization and Method

8

24. Lenz, R.; Enders, K.; Stedmon, C. A.; Mackenzie, D. M. A.; Nielsen, T. G., A

critical assessment of visual identification of marine microplastic using Raman

spectroscopy for analysis improvement. Mar. Pollut. Bull. 2015, 100 (1), 82-91.

25. Collard, F.; Gilbert, B.; Eppe, G.; Parmentier, E.; Das, K., Detection of

Anthropogenic Particles in Fish Stomachs: An Isolation Method Adapted to Identification

by Raman Spectroscopy. Arch. Environ. Contam. Toxicol. 2015, 69 (3), 331-9.

26. Pilat, Z.; Damborsky, J.; Prokop, Z.; Jezek, J.; Kizovsky, M.; Klementova, T.;

Kratky, S.; Sobota, J.; Samek, O.; Zemanek, P.; Buryska, T., Surface-enhanced Raman

Spectroscopy in Microfluidic Chips for Directed Evolution of Enzymes and

Environmental Monitoring. In 2019 PhotonIcs & Electromagnetics Research Symposium

- Spring (PIERS-Spring), 2019; pp 1301-1309.

27. Chung, K.-T., Azo dyes and human health: A review. J. Environ. Sci. Heal. C

2016, 34 (4), 233-261.

28. Za’bah, N. F.; Ralib, A. A. M.; Kwa, K. S. K.; O’Neill, A., Material

Characterization of a Doped Triangular Silicon Nanowire Using Raman Spectroscopy.

Adv. Sci. Lett. 2018, 24 (11), 8962-8965.

29. He, X.; Liu, X.; Nie, B.; Song, D., FTIR and Raman spectroscopy

characterization of functional groups in various rank coals. Fuel 2017, 206, 555-563.

30. Taz, H.; Ruther, R.; Malasi, A.; Yadavali, S.; Carr, C.; Nanda, J.;

Kalyanaraman, R., In Situ Localized Surface Plasmon Resonance (LSPR) Spectroscopy

to Investigate Kinetics of Chemical Bath Deposition of CdS Thin Films. J. Phys. Chem.

C 2015, 119 (9), 5033-5039.

Page 22: Raman Spectroscopy Substrate Optimization and Method

9

31. Li, L.; Steiner, U.; Mahajan, S., Single nanoparticle SERS probes of ion

intercalation in metal-oxide electrodes. Nano Lett. 2014, 14 (2), 495-8.

32. Berweger, S.; Neacsu, C. C.; Mao, Y.; Zhou, H.; Wong, S. S.; Raschke, M. B.,

Optical nanocrystallography with tip-enhanced phonon Raman spectroscopy. Nat.

Nanotechnol 2009, 4 (8), 496-9.

33. Huefner, A.; Kuan, W.-L.; Müller, K. H.; Skepper, J. N.; Barker, R. A.;

Mahajan, S., Characterization and Visualization of Vesicles in the Endo-Lysosomal

Pathway with Surface-Enhanced Raman Spectroscopy and Chemometrics. ACS Nano

2015, 10 (1), 307-316.

Page 23: Raman Spectroscopy Substrate Optimization and Method

10

Chapter 2

Raman Theory and Instrumentation

Page 24: Raman Spectroscopy Substrate Optimization and Method

11

2.1 Raman Spectroscopy

During the 1920s, significant interest was placed on explanation of scattering

phenomena by charged particles. The discovery of the Compton effect, by which x-ray

photons change energy when scattered by electrons, paved a theoretical way for

understanding of inelastic light scattering, which formed the basis for what C. V. Raman

(since knighted as Sir C. V. Raman) observed in the late 1920s as a “feeble flouorescence”

or “new radiation.”1The observed effect, called the Raman effect, is another light scattering

technique in which a small fraction of photons change frequency upon interaction with

matter. Initial Raman spectra were recorded using sunlight and photographic plates as the

excitation source and recording device, respectively, and these spectra could take 24 – 48

hours to record.2 With such long collection times and large volumes of liquid samples

required, Raman spectroscopy remained a niche technique until the 1960s, when lasers

became widely available for use as excitation sources.2 As instrumentation advanced, so

did the applicability of Raman as an analytical technique.

When light interacts with matter, the incident photons can undergo one of several

processes: absorption, transmittance, or scattering. For scattering processes, there are two

distinct outcomes from an energetic standpoint. If no change in energy occurs during the

scattering of the output photon, the process is known as elastic scattering. For scattering

processes where there is an energy difference in the scattered photons due to energy

transfer, these processes are referred to as inelastic scattering. These inelastic scattering

phenomena are collectively referred to as Raman scattering (Figure 2.1), and have two

Page 25: Raman Spectroscopy Substrate Optimization and Method

12

Figure 2.1. Jablonski diagram for Rayleigh and Raman scattering. E0 is the electronic ground state energy, E1 is the first excited electronic state energy, h is Planck’s constant, and ν0 and νm are the frequencies of the ground vibrational state and a vibrational mode, respectively.

Page 26: Raman Spectroscopy Substrate Optimization and Method

13

variations: the photon absorbs energy from a molecular vibration, and is scattered at a

higher energy (anti-Stokes scattering), and conversely, when the molecular vibration

absorbs energy from the photon, and the photon is scattered at a lower energy (Stokes

scattering).

2.2 Raman Theory An electrical dipole P is induced when a molecule is placed into an external electric

field E, and the relation between these quantities can be expressed as a power series

𝑷𝑷 = 𝛼𝛼 ∙ 𝑬𝑬 +12∙ 𝛽𝛽 ∙ 𝑬𝑬2 +

16∙ 𝛾𝛾 ∙ 𝑬𝑬3 + ⋯ 2.1

where α is the polarizability (~10-40 C·V-1·m2), β is the hyperpolarizability (~10-50 C·V-

2·m3), and γ is the 2nd hyperpolarizability (~10-60 C·V-3·m4).2 Since each successive

hyperpolarizability tensor decreases by a factor of 10-10, the contribution of all the

hyperpolarizability tensors can, to a first approximation, be ignored (equation 2.2), since

their contribution is minimal.

𝑷𝑷 = 𝛼𝛼 ∙ 𝑬𝑬 2.2

The polarizability (α) of a molecule is dependent on the nuclear coordinate (q) of

the molecule, and if a molecule is vibrating with a frequency νm, the nuclear displacement

is given by equation 2.3.

𝑞𝑞 = 𝑞𝑞0 cos 2𝜋𝜋𝜈𝜈𝑚𝑚𝑡𝑡 2.3

For small amplitude vibrations, α is a linear function of q,3 which can be written as

follows:

𝛼𝛼 = 𝛼𝛼0 + �𝜕𝜕𝛼𝛼𝜕𝜕𝑞𝑞�0∙ 𝑞𝑞 + ⋯ 2.4

Page 27: Raman Spectroscopy Substrate Optimization and Method

14

Substituting equation 2.3 into 2.4 and neglecting the higher order terms yields

𝛼𝛼 = 𝛼𝛼0 + �𝜕𝜕𝛼𝛼𝜕𝜕𝑞𝑞�0∙ 𝑞𝑞0 cos 2𝜋𝜋 𝜈𝜈𝑚𝑚𝑡𝑡 2.5

Light can be viewed as an oscillating electric field, and as such the electric field

vector E can be described as a function of time by equation 2.6, below.

𝑬𝑬 = 𝑬𝑬0cos 2𝜋𝜋𝜈𝜈0𝑡𝑡 2.6

Substituting equations 2.5 and 2.6 into equation 2.2 yields:

𝑷𝑷 = 𝛼𝛼 ∙ 𝑬𝑬 = �𝛼𝛼0 + �𝜕𝜕𝛼𝛼𝜕𝜕𝑞𝑞�0𝑞𝑞0 cos 2𝜋𝜋 𝜈𝜈𝑚𝑚𝑡𝑡� (𝑬𝑬0 cos 2𝜋𝜋𝜈𝜈0𝑡𝑡)

= 𝛼𝛼0𝑬𝑬0 cos 2𝜋𝜋𝜈𝜈0𝑡𝑡 + �𝜕𝜕𝛼𝛼𝜕𝜕𝑞𝑞�0𝑞𝑞0𝑬𝑬0 cos 2𝜋𝜋𝜈𝜈𝑚𝑚𝑡𝑡 cos 2𝜋𝜋𝜈𝜈0𝑡𝑡

2.7

Lastly, application of the trigonometric identity cos𝛼𝛼 cos𝛽𝛽 = 12

[cos(𝛼𝛼 + 𝛽𝛽) +

cos(𝛼𝛼 − 𝛽𝛽) yields

𝑷𝑷 = 𝛼𝛼0𝑬𝑬0 cos 2𝜋𝜋𝜈𝜈0𝑡𝑡 +

12�𝜕𝜕𝛼𝛼𝜕𝜕𝑞𝑞�0𝑞𝑞0𝑬𝑬0[cos(2𝜋𝜋(𝜈𝜈0 + 𝜈𝜈𝑚𝑚)𝑡𝑡)

+ cos(2𝜋𝜋 (𝜈𝜈0 − 𝜈𝜈𝑚𝑚)𝑡𝑡)]

2.8

The first term in equation 2.8 represents Rayleigh scattering, and the second term

represents anti-Stokes (energy increase) and Stokes scattering (energy decrease),

respectively.

The energy of the scattered photon in Raman spectroscopy does not coincide with

the energy difference between two electronic states, but instead is an intermediate value

between the electronic quantized electronic energy levels. During a Raman scattering

process, the molecule is excited to a “virtual state” that is less than the first excited

Page 28: Raman Spectroscopy Substrate Optimization and Method

15

electronic state (Figure 2.1). For anti-Stokes processes to occur, the molecule must already

be in a vibrationally excited state (ν ≥ 1) prior to interaction with the monochromatic light

source. This excitation is typically provided through thermal energy exchange, but the

majority of molecules remain in the vibrational ground state due to the Boltzman

distribution. This fundamental concept explains why the intensity of Stokes scattering is

greater than that of anti-Stokes scattering, and due to this it is most common to measure

the Stokes scattered photons when conducting Raman experiments (Aroca citation).

The resulting Raman spectrum that is obtained from a measurement is shown as a

plot of Raman intensity as a function of Raman shift, which has units of wavenumbers (cm-

1). A representative spectrum can be found in Figure 2.2. The Raman shift is a relative

measure of the energy of a molecular vibration, since the energy difference from the input

photon corresponds to the vibrational energy level difference. Since these measures are

relative, they are independent of the incident laser frequency. The Raman wavenumber (ω)

of a vibration can be calculated by using Equation 2.9, below.

𝜔𝜔 = 𝜐𝜐𝑚𝑚� − 𝜐𝜐0� =𝜐𝜐𝑚𝑚𝑐𝑐−𝜐𝜐0𝑐𝑐

2.9

Raman signal is inherently weak, due in large part to the small Raman scattering

cross-sections (σRS) of most molecules. The total Raman scattered light (IRS) is given by

Equation 2.10

𝐼𝐼𝑅𝑅𝑅𝑅 = 𝜎𝜎𝑅𝑅𝑅𝑅𝐼𝐼0 2.10

Page 29: Raman Spectroscopy Substrate Optimization and Method

16

Figure 2.2. A typical Raman spectrum obtained from neat pyridine, a strong Raman scatterer. The spectrum is plotted as a with intensity as a function of Raman shift, and was obtained in 1 s using 633 nm excitation at 5.7 mW of power.

Page 30: Raman Spectroscopy Substrate Optimization and Method

17

where I0 is the incident photon flux (photons cm-2 s-1).4

Raman cross sections are

proportional to the square of the polarizability derivative for the transition, and the fourth

power of the scattering frequency ωs, such that

𝜎𝜎𝑅𝑅𝑅𝑅 ∝ 𝜔𝜔𝑠𝑠4|𝛼𝛼ʹ1→2|2 2.11

Physically, the Raman cross-section represents the total area presented by a molecule for a

scattering process. Typical values for σRS are on the order of 10-30 to 10-25 cm2 molecule-1

under normal Raman scattering conditions.4, 5

In quantum mechanics, a transition from state 1 to state 2 will only occur if the

transition moment integral (equation 2.12) between the two states is nonzero.6

𝜇𝜇1→2 = �𝛹𝛹2∗𝑃𝑃𝛹𝛹1𝑑𝑑𝑑𝑑 2.12

The transition moment integral forms the basis for spectroscopic selection rules. Since

Raman is an inelastic scattering technique, we are concerned primarily with the second

term in equation 2.8, which can be nonzero under the following conditions: 1) There must

be a change in polarizability during a vibration for the vibration to be Raman active, and

2) only transitions between adjacent vibrational levels are allowed (Δν = ±1)

Using equation 2.2 and the fact that the induced dipole in a molecule is a three-

dimensional vector, and can be represented in matrix form by

�𝑃𝑃𝑥𝑥𝑃𝑃𝑦𝑦𝑃𝑃𝑧𝑧� = �

𝛼𝛼𝑥𝑥𝑥𝑥 𝛼𝛼𝑥𝑥𝑦𝑦 𝛼𝛼𝑥𝑥𝑧𝑧𝛼𝛼𝑦𝑦𝑥𝑥 𝛼𝛼𝑦𝑦𝑦𝑦 𝛼𝛼𝑦𝑦𝑧𝑧𝛼𝛼𝑧𝑧𝑥𝑥 𝛼𝛼𝑧𝑧𝑦𝑦 𝛼𝛼𝑧𝑧𝑧𝑧

� �𝐸𝐸𝑥𝑥𝐸𝐸𝑦𝑦𝐸𝐸𝑧𝑧� 2.13

Page 31: Raman Spectroscopy Substrate Optimization and Method

18

leads to the identification of the polarizability tensor (the 3 x 3 matrix in equation 2.13).

This tensor is symmetric (αxy = αyx, etc.) in normal Raman scattering, and so long as one

of the tensor components is nonzero, the vibration will be Raman-active.3

2.3 Surface-Enhanced Raman Spectroscopy

Raman spectroscopy is capable of providing a wealth of information about the

structure of the analyte molecule, and the use of multivariate analysis has allowed for the

characterization of complex systems, including biological materials.5 However, due to

extremely small scattering cross-sections (10-30 to 10-25 cm2 molecule-1) which require a

significant number of analyte molecules to be present before signal becomes noticeable,

normal Raman spectroscopy is less useful for ultrasensitive analysis.5

In 1974, Fleischman et al. observed greatly enhanced pyridine signal after

adsorption to a roughened silver electrode. The authors postulated that the enhancement in

the signal was due to increasing the surface area of the electrode through the roughening

process, which in turn allowed for more molecules of pyridine to adsorb onto the surface.7

Jeanmaire and Van Duyne8 proposed that this enhancement effect was the result of

increased electric fields, whereas Albrecht and Creighton9 hypothesized that the anomalous

results were the result of resonance Raman scattering from molecular electronic states.7

From equation 2.2, these correspond to altering of the polarizability (chemical

enhancement) or altering the electric field (electric field enhancement). While initially

there was some conflict as to which enhancement mechanism was responsible for the

improvement in signal, it is now largely accepted that the primary contributor is the

electromagnetic enhancement mechanism proposed by Jeanmaire and Van Duyne.8, 10

Page 32: Raman Spectroscopy Substrate Optimization and Method

19

By placing molecules of interest into proximity of a plasmonic metal nanostructure, an

enhanced electric field is produced. Upon interaction with the excitation laser, the

conduction band electrons start to oscillate. These collective oscillations are referred to as

the localized surface plasmon resonance (LSPR) shown in Figure 2.3. The resultant electric

field produced by the LSPR of the metal enhances the Raman scattering of an analyte

molecule, in a technique that is referred to as surface-enhanced Raman scattering (SERS).

The SERS cross section shows a marked increase over normal Raman, with typical SERS

cross sections on the order of 10-16 cm2 molecule-1, which is an increase of 109 - 1014 over

normal Raman cross sections.11 Excitation of the LSPR has two consequences: selective

absorption and scattering of the resonant electromagnetic radiation, and generation of large

magnetic fields at the surface of the roughened feature.12 Concentration of the incident light

occurs preferentially in gaps, crevices, or sharp features of plasmonic materials,10 and

theoretical calculations of enhanced electric fields on nanoparticles can be seen in Figure

2.4.

Enhancement from SERS substrates is heightened due to the presence of “hot

spots,” which are found in the area between two adjacent nanostructures where the electric

field magnitude is greatest.13 These can be seen in Figure 2.4 as the red areas on the surface

of each nanoparticle morphology. Because this electric field is localized at the surface and

in the hot spots between substrates, SERS enhancement demonstrates a strong distance

dependence where the most intense signal is obtained at the surface and when the molecule

Page 33: Raman Spectroscopy Substrate Optimization and Method

20

Figure 2.3. The localized surface plasmon resonance (LSPR) phenomenon, in which conduction band electrons of a metal nanostructure oscillate in the presence of an external electric field.

Page 34: Raman Spectroscopy Substrate Optimization and Method

21

Figure 2.4. Theoretical simulations of the EM field enhancement around silver nanoparticles, with single triangular (top left) and ellipsoidal morphologies (right), as well as dimers of spherical nanoparticles (bottom left), shown.12

Page 35: Raman Spectroscopy Substrate Optimization and Method

22

is adsorbed within a distance of 2 nm.13-15 The electromagnetic (EM) enhancement of

SERS does not require the analyte molecule to be in direct contact with the enhancing

surface, but within a certain sensing volume.15, 16 However, this enhancement is a distance-

dependent phenomenon. Since the electric field decays around a spherical particle, which

is the simplest example of an enhancing substrate, proportional to d-3, and EM enhancement

can be approximated as the fourth power of the electric field (|E|4),14-16 the EM

enhancement would appear to decay proportional to d-12.4, 16 When the surface area scaling

(which is proportional to d2) is considered, the overall distance dependence shows a d-10

proportionality.14-16 For most substrates, the SERS intensity drops off sharply within a

couple nanometers of the enhancing surface, and silver-film-over-nanosphere (AgFON)

substrates have been shown to have SERS intensity reduced to 7-10% of the original

intensity at a distance of 3 nm from the substrate surface.14, 15

Common SERS substrate materials include the noble/coinage metals Ag, Au, and

Cu. For a material to exhibit plasmonic properties, two criteria must be met with regards

to the complex dielectric function of the metal: the real component of the dielectric constant

should be large and negative, and the imaginary component of the dielectric constant

should be small and positive.10 For the above metals, these conditions are met in the visible

and near-IR regions of the electromagnetic spectrum, which makes them ideal SERS

substrates for use in these regions. A wide range of SERS substrates exist, from

nanoparticles of simple (spheres,17 cubes, prisms, rods, and octahedra18) to complex

(nanorice,19, 20nanourchin,21) morphologies. In addition to these, 2D SERS substrates have

seen an increase in fabrication due to their reproducible enhancement. Most notable among

these 2D substrates are film-over-nanosphere (FON) substrates, which are created using

Page 36: Raman Spectroscopy Substrate Optimization and Method

23

chemical vapor deposition (CVD) to lay a thin film of silver or gold over polystyrene or

silica spheres.10

Extensive theoretical modeling has been performed to provide insight into

electromagnetic enhancement by noble-metal (Ag, Au, and Cu) nanoparticles,22, 23 and the

simple example of an isolated sphere with a quasistatic description of the incident

electromagnetic field has found that the proportionality shown in Equation 2.13.

𝐸𝐸2 ∝ 𝐸𝐸02 �𝜀𝜀𝑚𝑚 − 𝜀𝜀0𝜀𝜀𝑚𝑚 − 2𝜀𝜀0

�2 2.13

In this equation, E is the electric field at the surface of the sphere, E0 is the incident field

magnitude, εm is the wavelength-dependent dielectric constant of the nanoparticle

comprising metal, and ε0 is the local environment dielectric constant12. This proportionality

demonstrates that when εm = -2ε0, the electric field at the surface of the sphere becomes

greatly enhanced12. For the noble metals, this condition can be met with excitation

wavelengths in the visible or near-infrared regime of the electromagnetic spectrum, and it

has been demonstrated that these noble metals are SERS active in the wavelength ranges

shown in Figure 2.5.

For spheres, the phenomena of scattering and extinction are wholly explained by

analytical solutions to Maxwell’s equations (Mie Theory).4 These processes are

proportional to a g factor, which is defined in Equation 2.13:

Page 37: Raman Spectroscopy Substrate Optimization and Method

24

Figure 2.5. Approximate wavelength ranges where Ag, Au, and Cu have been well-characterized and are established to support SERS.10

Page 38: Raman Spectroscopy Substrate Optimization and Method

25

𝑔𝑔 =𝜀𝜀(𝜔𝜔) − 𝜀𝜀𝑚𝑚𝜀𝜀(𝜔𝜔) + 2𝜀𝜀𝑚𝑚

2.13

where ε(ω) is the dielectric function of the metal, and εm is the dielectric constant of the

medium/environment.4 The plasmon resonance condition for spheres is met when ε(ω) = -

2εm. Similarly, for a cylinder, the g-factor is

𝑔𝑔 =𝜀𝜀(𝜔𝜔) − 𝜀𝜀𝑚𝑚𝜀𝜀(𝜔𝜔) + 𝜀𝜀𝑚𝑚

2.14

and the plasmon resonance condition for the cylinder is met when ε(ω) = -εm.24 It is worth

noting that since ε(ω) is a complex number, it will never exactly equal εm. However, when

the real part of ε(ω) meets the resonance condition, then the system will demonstrate a

large response and local field enhancement will be increased.24 The magnitude of the

imaginary component of the complex dielectric constant of the metal governs the intensity

of the heightened response.

SERS has several applications in a variety of fields, including polymer and materials

science, biochemistry and biosensing, catalysis, and electrochemistry.10 The Van Duyne

group has made significant progress in development of a glucose biosensor, using a silver

film over nanosphere SERS sensor along with spatially offset Raman spectroscopy

(another variant of Raman not discussed in this current work) to detect in vivo glucose

concentrations that are below the currently accepted lower limit established by the

International Organization Standard requirements,10 which shows considerable promise for

those living with diabetes mellitus. Additionally, SERS has been used in the detection of

various other diseases, including cancers, Alzheimer’s disease, and Parkinson’s disease.10

Non-biological applications of SERS include the detection of explosives (e.g. half-mustard

Page 39: Raman Spectroscopy Substrate Optimization and Method

26

agent and dinitrobenzenthiol), as well as spectroelectochemical applications regarding the

behavior of electrochemically active molecules in differing oxidation states via

spectroscopy.10

2.4 Enhancing Substrate Synthesis and Fabrication

As mentioned above, the most commonly used plasmonic metals for SERS

substrates are the coinage/noble metals gold (Au) and silver (Ag). These metals can be

synthesized into a variety of morphologies through simple to complex wet chemical

methods. Additionally, two-dimensional (2D) materials have been employed to provide

more reproducible SERS enhancement. For the projects and data shown in this dissertation,

colloidal Au and Ag nanoparticles have been employed, in addition to 2D substrates. The

synthesis/fabrication of these substrates is explored in the subsections below.

2.4.1 Gold Nanoparticle Synthesis

The Frens25 synthesis was commonly used to generate gold colloids capable of

SERS enhancement. Briefly, a 45 mL volume of 0.01% (w/w) hydrogen

tetrachloroaurate(III) (HAuCl4) was prepared in ultrapure deionized (>18 MΩ cm) water

and brought to a rapid boil while magnetically stirred. To this solution, 300 μL of 1% (w/w)

trisodium citrate dihydrate (Na3C6H5O7 ·2 H2O) was added rapidly. The resulting solution

was allowed to boil for 5 minutes, over the course of which the color changed from a pale

yellow to dark blue to wine red. The resulting colloids were characterized using scanning

electron microscopy (SEM) and UV-visible extinction spectroscopy (Agilent Cary 5000

UV/Vis/NIR). A typical SEM image and extinction spectrum are found in Figure 2.6.

Page 40: Raman Spectroscopy Substrate Optimization and Method

27

Figure 2.6. SEM image (a) and extinction spectrum (b) of colloidal gold synthesized by the Frens method.

Page 41: Raman Spectroscopy Substrate Optimization and Method

28

2.4.2 Silver Nanoparticle Synthesis

Colloidal silver nanoparticles were using a modification of the synthesis described

by Leopold and Lendl.26 Briefly, 150 μL of 1 M hydroxylamine hydrochloride (NH3OH

·HCl, HA), 300 μL of 1 M sodium hydroxide (NaOH), and 89.550 mL of ultrapure

deionized water were combined under magnetic stirring (300 rpm), yielding final

concentrations of 1.67 mM and 3.33 mM HA and NaOH, respectively. To this solution, 10

mL of 10 mM silver nitrate (AgNO3) was rapidly added, resulting in a brown-green

solution of colloidal hydroxylamine-reduced silver nanoparticles (HAAgNP). The

resulting colloids were characterized using SEM and UV-visible extinction spectroscopy.

An SEM image and extinction spectrum can be found in Chapter 3 of this work.

2.4.3 2D/3D SERS Substrates

Historically, the enhancing substrates for SERS measurements started with

roughened electrode work by Fleishman et al., and moving to colloidal suspensions of gold

and silver.27 Because of size distributions obtained in synthesis of colloidal gold and silver

nanoparticles, interest has shifted toward 2- and 3-dimensional substrates for SERS

enhancement, including metal films of various thicknesses and nanostructured features,

and recent advances in the field have moved to hybrid techniques of both colloids and

films.27 Rigorous characterization of SERS substrates is required (and is often far too

complex with colloidal substrates) to extract analytical parameters, including sensitivity

and reproducibility. By introducing these multi-dimensional systems into the SERS

conversation, it is possible to have highly enhancing reproducible substrates that can be

employed in analytical settings.

Page 42: Raman Spectroscopy Substrate Optimization and Method

29

Substrates including aluminum,28 silver,29 and gold30 film-over-nanosphere (FON)

substrates have been employed in SERS-based detection schemes in recent times. FON

substrates consist of a close-packed array of silica, polystyrene, or latex microspheres that

are deposited on a silicon wafer or glass substrate prior to film deposition of the appropriate

plasmonic metal over the microspheres. The close packing of the spheres provides crevices

in between the spheres, which then generate a hot spot once the plasmonic film is deposited.

Zinc oxide nanowires have been deposited with silver nanoparticles and used to detect

amoxicillin, an antibacterial agent that is used for human and zoonotic diseases that has

been found in foodstuffs, at the part-per-billion concentration range.31 Silver “nanocactus”

substrates, consisting of silver dendrites decorated with AgNP-tipped Si nanoneedles have

been used to detect malachite green, an antifungal/antiparasitic that also has carcinogenic,

mutagenic, and toxic properties, in water supplies, resulting in detection limits in the

picomolar (pM) range.32 Similarly, Si micropillar arrays have been decorated with

germanium nanotapers with AgNP located at the point of the taper to provide a well-

enhancing SERS substrate capable of detecting toxic organic pollutants in the micromolar

concentration range.33 All the aforementioned substrates have shown enhancement factors

of >107 (vide infra, Section 2.5).

Though EM enhancement is the preferred method for amplifying Raman signal,

chemical enhancement mechanisms also exist. Two-dimensional transition metal

dichalcogenides have been explored as chemically enhancing substrates through the use of

2H and 1T phases of molybdenum disulfide and diselenide (MoS2 and MoSe2,

respectively).34 Each phase was prepared as a nanosheet supported on a Si/SiO2 substrate

and then exposed to rhodamine 6G and crystal violet, both strongly Raman scattering

Page 43: Raman Spectroscopy Substrate Optimization and Method

30

molecules, at decreasing concentrations. With these strongly scattering reporter molecules,

the dichalcogenide substrates were found to have similar detection limits.

In this dissertation, a disc-on-pillar array was investigated as a sensing medium for

volatile organic compounds. The fabrication of the arrays was carried out using established

methods by Agapov et al.,35 in which a Si wafer was deposited with a platinum film then

thermally annealed to generate circular Pt metal islands on the surface of the wafer. The

wafer was then reactive ion etched to generate a pillar morphology, followed by deposition

of an Au film, resulting in plasmonic metal discs on the top of the pillars. Previous work

has characterized this substrate with a high enhancement factor (~107), allowing it to be

used for sensitive detection. To facilitate gas capture at the plasmonic surface, a porous

SiO2 layer was deposited on the substrate, providing a convoluted network for the

molecules to traverse before escaping the enhancing surface.

2.5 Enhancement Factors

There are three predominant methods to calculate enhancement factors (EF) for

SERS measurements: single-molecule enhancement factor (SMEF), SERS substrate

enhancement factor (SSEF), and analytical enhancement factor (AEF).36 These first two

EF calculations focus on intrinsic properties of the substrate itself. In many instances,

though, one is simply concerned with the amount of signal that can be observed under

experimental conditions. Analytical enhancement factors (AEFs) provide an intuitive and

relevant alternative. The AEF is defined as:

𝐴𝐴𝐸𝐸𝐴𝐴 = 𝑑𝑑𝜎𝜎𝑅𝑅𝑆𝑆𝑅𝑅𝑅𝑅 𝑑𝑑𝑑𝑑⁄𝑑𝑑𝜎𝜎𝑅𝑅𝑎𝑎𝑚𝑚𝑎𝑎𝑎𝑎 𝑑𝑑𝑑𝑑⁄ ∙

𝑡𝑡𝑎𝑎,𝑅𝑅𝑎𝑎𝑚𝑚𝑎𝑎𝑎𝑎

𝑡𝑡𝑎𝑎,𝑅𝑅𝑆𝑆𝑅𝑅𝑅𝑅 2.15

Page 44: Raman Spectroscopy Substrate Optimization and Method

31

where dσ/dΩ is the differential Raman cross sections (for SERS or normal Raman

scattering, as indicated by the subscript), and ta is the acquisition time for each spectrum.37

The differential Raman cross section can be solved numerically by use of equation 2.16

𝑑𝑑𝜎𝜎𝑅𝑅𝑆𝑆𝑅𝑅𝑅𝑅 𝑑𝑑𝑑𝑑⁄𝑑𝑑𝜎𝜎𝑅𝑅𝑎𝑎𝑚𝑚𝑎𝑎𝑎𝑎 𝑑𝑑𝑑𝑑⁄ =

𝐼𝐼𝐴𝐴𝑅𝑅𝑆𝑆𝑅𝑅𝑅𝑅 𝑐𝑐𝑅𝑅𝑆𝑆𝑅𝑅𝑅𝑅⁄𝐼𝐼𝐴𝐴𝑅𝑅𝑎𝑎𝑚𝑚𝑎𝑎𝑎𝑎 𝑐𝑐𝑅𝑅𝑎𝑎𝑚𝑚𝑎𝑎𝑎𝑎⁄ 2.16

where IA is the integrated area under a specific peak in the SERS or normal Raman

spectrum, and c is the concentration of the sample analyzed by either SERS or normal

Raman. Substitution of 2.16 into 2.15 followed by rearrangement yields

𝐴𝐴𝐸𝐸𝐴𝐴 = 𝐼𝐼𝐴𝐴𝑅𝑅𝑆𝑆𝑅𝑅𝑅𝑅𝐼𝐼𝐴𝐴𝑅𝑅𝑎𝑎𝑚𝑚𝑎𝑎𝑎𝑎

𝑐𝑐𝑅𝑅𝑎𝑎𝑚𝑚𝑎𝑎𝑎𝑎𝑐𝑐𝑅𝑅𝑆𝑆𝑅𝑅𝑅𝑅

𝑡𝑡𝑎𝑎,𝑅𝑅𝑎𝑎𝑚𝑚𝑎𝑎𝑎𝑎

𝑡𝑡𝑎𝑎,𝑅𝑅𝑆𝑆𝑅𝑅𝑅𝑅 2.17

2.6 Raman Instrumentation

A typical Raman spectrometer consists of four components: an excitation source, a

sample, a dispersion system, and a detector. Light that is emitted from a laser must also be

efficiently delivered to the sample for excitation, and then the Raman scattered light must

be efficiently directed to the dispersion system and detector. These processes are directed

by selection of the appropriate optical components in the system. These systems can be

provided commercially as all-in-one box systems or can be home built in a research lab.

All of the experiments described within this dissertation were carried out on a home-built

Raman system.

2.6.1 Excitation Sources

Prior to the advent of highly monochromatic excitation sources provided by lasers,

the typical excitation source for Raman experiments was a mercury arc lamp, which has a

Page 45: Raman Spectroscopy Substrate Optimization and Method

32

several emission lines that range from 435 – 579 nm.38 Due to the polychromatic nature of

this light, it was necessary that early Raman systems be equipped with a series of filters

that were employed to isolate a specific wavelength or broadband series of wavelengths.

However, these lamps had many disadvantages up to and including failure of the lamp to

ignite. The development of lasers made Raman spectroscopy measurements surpassingly

easier, due to improved frequency stability, narrow bandwidths (typically below 1 cm-1),

the absence of side bands, and low divergence.2 Lasers also are able to be tightly focused,

which allows for the minimization of sample volumes in Raman scattering experiments.

The initial lasers that were used consisted of gas lasers that provided light in the visible

region of the electromagnetic spectrum. Modern lasers are predominantly continuous wave

(CW) lasers, which allow illumination to be consistent at the sample, and requires lower

powers. The majority of Raman experimentation is conducted using these CW lasers, with

the most common systems involving the use of either gas lasers (e.g. argon or krypton),1

or diode lasers. Diode lasers provide numerous advantages over traditional gas lasers,

including reduced spatial footprints, longer lifetimes, and lower cost, with near-IR diode

lasers providing greater separation between the Raman shifted output and the native sample

fluorescence.39

The data contained in this dissertation was all collected using a series of visible and

near-infrared (IR) lasers. Two optically pumped semiconductor lasers were used, with

output wavelengths of 488 nm and 532 nm (Coherent Technologies). A helium-neon laser

(He-Ne) was employed for excitation at 633 nm, and a diode laser was used with an output

of 785 nm in the near-IR.

Page 46: Raman Spectroscopy Substrate Optimization and Method

33

2.6.2 Samples and Collection Optics

One of the key advantages of Raman spectroscopy lies in the fact that samples can

be analyzed in any physical state of matter – gas, liquid, solution, or solid. This versatility

is highlighted throughout this dissertation, since the materials analyzed within fall into all

of the above states.

The excitation light source is directed toward the sample using optics and

optomechanics to manipulate the beam path toward the sample. Typically, the laser line is

aligned using a system that is comprised of 2 mirrors and 2 irises, which aligns the laser

both in the plane of the laser table and vertically, assuring that there is no deviation of the

laser line from parallel to the table. In macroscopic Raman systems, the light is focused

onto the sample using a focusing lens that is matched to the focal length between the lens

and the surface of the sample. In microscopic systems, the excitation light source is directed

into a microscope, where the microscope objective serves to focus the light onto the

sample. In both cases, focus of the light onto the sample assures that a tight region of the

sample is irradiated, which raises the power density of the laser spot.

The scattered light must also be effectively directed to the dispersion system, so

optics are employed here to facilitate this. In both macro- and microscopic systems, this is

achieved through the use of collection lenses, which are typically have large diameters to

ensure that the majority of the scattered light is collected by the lens. The collection lens

serves to focus the light onto the entrance slit of the dispersion system, or spectrometer.

The collection power of a lens is specified by the lens F-number (f/#), which is defined

mathematically as shown in Equation 2.18, where f is the focal length of the lens and D is

Page 47: Raman Spectroscopy Substrate Optimization and Method

34

the lens diameter; it is of utmost importance that the f/# of the collecting lens match the f/#

of the spectrometer.3

𝑓𝑓/# =𝑓𝑓𝐷𝐷

2.18

Most Raman systems also include a movable sample stage, which can be as simple as an

XYZ translation stage equipped with micrometers or can be fully motorized piezoelectric

stages that can be adjusted in micron intervals.

Prior to entry into the dispersion system, it is necessary to remove the Rayleigh

scattered light so that the Raman signal is not suppressed by the intensity of the Rayleigh

line. This is typically achieved using a holographic notch filter, which has a high optical

density and exhibits a sinusoidal transmission profile, which removes the possibility of

reflection bands. For Stokes shifted Raman photons, the wavelength is longer than the

Rayleigh wavelength, which means that long pass filters are used in typical systems, and

these filters have extremely narrow bandwidths.

2.6.3 Dispersion System

The dispersion system utilized in Raman instrumentation is commonly referred to

as a spectrograph or spectrometer. The purpose of the instrument is to disperse the light

that enters the spectrometer into the component wavelengths (or specifically, Raman

shifted photons). The dispersive element in a spectrometer is the medium that provides this

spatial separation of the Raman scattered light. Prisms are an example of a dispersive

element, but these are less common in modern Raman spectrometers. The current

technology relies on the use of diffraction gratings, which is a linear repetition of reflecting

elements, which are commonly referred to as grooves.2 The dispersive behavior of the

Page 48: Raman Spectroscopy Substrate Optimization and Method

35

grating can be described by the grating equation (Equation 2.19), where θi and θm are the

angles between the normal of the grating surface and the incident and diffracted light,

respectively, λ is the wavelength, d is the distance between the grooves on the grating, and

m is the integer that describes the order of the diffracted light.40, 41

𝑑𝑑(sin𝜃𝜃𝑚𝑚 − sin𝜃𝜃𝑖𝑖) = 𝑚𝑚𝑚𝑚 2.19

Noting that 1/d is defined as the grating frequency, which is typically expressed in

grooves/mm, Equation 2.19 can be rewritten as Equation 2.20, where N is the grating

frequency.

sin𝜃𝜃𝑚𝑚 − sin 𝜃𝜃𝑖𝑖 =𝑚𝑚𝑚𝑚𝑑𝑑

= 𝑁𝑁𝑚𝑚𝑚𝑚 2.20

The grating frequency is responsible for the resolution of the grating, with higher grating

frequencies providing higher resolving power, while reducing the Raman shift range that

can be probed by the grating.

2.6.4 Detectors

As mentioned previously, Raman signal is inherently weak, and this weakness

creates an obstacle in detection schemes that must be overcome. Early Raman work was

conducted using photographic plates, which required exceedingly long development times

to obtain the spectral readout. This long time-frame made Raman signal acquisition less

desirable in laboratory settings until the advent of strong laser excitation sources and

sensitive detection techniques, including photon multiplier tubes (PMTs) and charge-

coupled devices (CCDs).

Page 49: Raman Spectroscopy Substrate Optimization and Method

36

PMTs consist of a photocathode that generates primary electrons when struck by a

photon. These primary electrons are directed to a series of dynodes, which generate

secondary electrons when impacted by a primary electron. This electron cascade continues

through the series of dynodes until the large packet of electrons reaches the anode, which

generates an electrical signal. The performance of a PMT is dictated by the quantum

efficiency of the PMT, which is wavelength dependent.3 PMTs suffer the disadvantage of

high background noise, or dark current, which is the outcome of spontaneous emission of

electrons from the photocathode and dynodes, but this can be somewhat mitigated by

thermoelectric cooling of the PMT. This can lead to condensation on the lens, which

requires that PMTs be periodically disassembled and dried.

Current Raman instrumentation makes use of CCD detectors, which are silicon-

based semiconductor arrays that generate electrons from incident photons and store the

signal as a small charge. The charge is stored in each pixel during the illumination, and at

the end of the illumination the signal is transferred to a serial register, which stores the

signal until readout. CCDs have the advantage of low readout noise, which occurs during

the movement of the signal from the serial register to the output device.3 CCDs have high

quantum efficiencies and sensitivity over a broad wavelength range, and as such need no

optical signal intensification.3

2.6.5 Raman Instrumentation in the Sharma Lab

The work contained within this dissertation was all performed on a home-built

confocal Raman microscope setup built in the Sharma Lab for use in the visible and near-

IR range. A schematic of this setup is found in Figure 2.7. The micro-Raman system

Page 50: Raman Spectroscopy Substrate Optimization and Method

37

consists of 4 excitation laser choices (488 nm, 532 nm, 633 nm, and 785 nm). All the laser

outputs are directed onto the same beam path, then are turned by a mirror to pass through

a set of irises. These irises are used to align the laser beam in the plane of the laser table,

while also ensuring that there are no vertical deviations in the beam path. This alignment

stage is critical for Raman spectroscopy, as it assures that the excitation light is provided

into the microscope setup with optimal intensity.

After alignment with the irises, the light goes through a series of focusing lenses

that correct for any spot expansion during the prior segments of the beam path. These lenses

also serve to focus the excitation light onto a set of periscope optics that raise the beam

path to the back of a Nikon Ti-u inverted microscope. Upon entry into the microscope, the

light is focused onto the sample using a microscope objective selected from the available

options of 20×, 40×, and 100× oil immersion objectives. When the excitation light finishes

is provided to the sample, the spot size is on the order of μm in diameter. The objective

further serves to collect the scattered light from the sample and sends it through an

internally mounted dichroic mirror to remove the Rayleigh scattered light. Upon exiting

the microscope, the light is collected by a 2ӯ planoconvex lens, which focuses the Raman

scattered light onto the entrance slit of a IsoPlane 320SCT spectrometer (Princeton

Instruments). Prior to the entrance slit, a holographic notch filter (Semrock) was placed in

the beam path to remove any residual Rayleigh scattered photons. All optics (Thorlabs) are

uncoated optics, which allows for use with a variety of excitation wavelengths.

The IsoPlane spectrometer is equipped with three different gratings mounted onto

a rotating turret. The gratings are 600, 1200, and 1800 grooves/mm, with the 600

groove/mm grating blazed at 750 nm, and the other two gratings blazed at 500 nm. The

Page 51: Raman Spectroscopy Substrate Optimization and Method

38

blaze provides the wavelength at which the grating is optimal, hence the 600 groove/mm

grating was used in experiments involving 785 nm and 633 nm excitation, while the 1200

and 1800 groove/mm gratings were used in experiments with 488 nm and 532 nm

excitation. After dispersion by the grating, the spatially separated Raman signal is

dispersed onto a PIXIS 400 CCD camera (Princeton Instruments).

Page 52: Raman Spectroscopy Substrate Optimization and Method

39

Figure 2.7. Schematic of the home-built Raman microscope system in the Sharma Lab, showing 4 excitation source options and indicating the beam path for each source.

Page 53: Raman Spectroscopy Substrate Optimization and Method

40

2.7 References

1. Grasselli, J. G.; Bulkin, B. J., Analytical Raman Spectroscopy. Wiley: Chichester,

1991.

2. Vandenabeele, P., Practical Raman Spectroscopy: An Introduction. Ghent

University: Belgium, 2013.

3. Ferraro, J. R.; Nakamoto, K.; Brown, C. W., Introductory Raman Spectroscopy.

2nd ed.; Academic Press: 2003.

4. Aroca, R., Surface-Enhanced Vibrational Spectroscopy. John Wiley & Sons, Ltd.:

2006.

5. Kneipp, K.; Kneipp, H.; Itzkan, I.; Dasari, R. R.; Feld, M. S., Ultrasensitive

chemical analysis by Raman spectroscopy. Chem. Rev. 1999, 99 (10), 2957-76.

6. McQuarrie, D. A., Quantum Chemistry. 2nd ed.; Mill Valley, California, 2008.

7. Campion, A.; Kambhampati, P., Surface-enhanced Raman scattering. Chem. Soc.

Rev. 1998, 27, 241-250.

8. Jeanmaire, D. L.; Van Duyne, R. P., Surface Raman Spectroelectrochemistry Part

I. Heterocyclic, Aromatic, and Aliphatic Amines Adsorbed on the Anodized Silver

Electrode. J. Electroanal. Chem. 1977, 8, 1-20.

9. Albrecht, M. G.; Creighton, J. A., Anomalously Intense Raman Spectra of

Pyridine at a Silver Electrode. J. Am. Chem. Soc. 1977, 99 (15), 5215-5217.

10. Sharma, B.; Frontiera, R. R.; Henry, A.-I.; Ringe, E.; Van Duyne, R. P., SERS:

Materials, applications, and the future. Mater. Today 2012, 15 (1-2), 16-25.

Page 54: Raman Spectroscopy Substrate Optimization and Method

41

11. Kneipp, K.; Kneipp, H.; Itzkan, I.; Dasari, R. R.; Feld, M. S., Surface-Enhanced

Raman Scattering and Biophysics. J. Phys.: Condens. Matter 2002, 14, R597-R624.

12. Haynes, C. L.; McFarland, A. D.; Van Duyne, R. P., Surface-enhanced Raman

spectroscopy. Anal. Chem. 2005, 77 (17), 338A-346A.

13. Moore, T. J.; Moody, A. S.; Payne, T. D.; Sarabia, G. M.; Daniel, A. R.;

Sharma, B., In Vitro and In Vivo SERS Biosensing for Disease Diagnosis. Biosensors

(Basel) 2018, 8 (2).

14. Dieringer, J. A.; McFarland, A. D.; Shah, N. C.; Stuart, D. A.; Whitney, A. V.;

Yonzon, C. R.; Young, M. A.; Zhang, X.; Van Duyne, R. P., Surface enhanced Raman

spectroscopy: new materials, concepts, characterization tools, and applications. Faraday

Discuss. 2006, 132, 9-26.

15. Masango, S. S.; Hackler, R. A.; Large, N.; Henry, A. I.; McAnally, M. O.;

Schatz, G. C.; Stair, P. C.; Van Duyne, R. P., High-Resolution Distance Dependence

Study of Surface-Enhanced Raman Scattering Enabled by Atomic Layer Deposition.

Nano Lett. 2016, 16 (7), 4251-9.

16. Stiles, P. L.; Dieringer, J. A.; Shah, N. C.; Van Duyne, R. P., Surface-enhanced

Raman spectroscopy. Annu. Rev. Anal. Chem. 2008, 1, 601-26.

17. Lin, H. X.; Li, J. M.; Liu, B. J.; Liu, D. Y.; Liu, J.; Terfort, A.; Xie, Z. X.;

Tian, Z. Q.; Ren, B., Uniform gold spherical particles for single-particle surface-

enhanced Raman spectroscopy. Phys. Chem. Chem. Phys. 2013, 15 (12), 4130-5.

18. Wang, Y.; Wang, E., Nanoparticle SERS Substrates. In Surface Enhanced Raman

Spectroscopy, Schlücker, S., Ed. Wiley-VCH: Weinheim, Germany, 2011; pp 39-69.

Page 55: Raman Spectroscopy Substrate Optimization and Method

42

19. Wang, H.; Brandl, D. W.; Le, F.; Nordlander, P.; Halas, N. J., Nanorice: a hybrid

plasmonic nanostructure. Nano Lett. 2006, 6 (4), 827-32.

20. Li, M.; Cushing, S. K.; Liang, H.; Suri, S.; Ma, D.; Wu, N., Plasmonic nanorice

antenna on triangle nanoarray for surface-enhanced Raman scattering detection of

hepatitis B virus DNA. Anal. Chem. 2013, 85 (4), 2072-8.

21. Liu, Z.; Cheng, L.; Zhang, L.; Jing, C.; Shi, X.; Yang, Z.; Long, Y.; Fang, J.,

Large-area fabrication of highly reproducible surface enhanced Raman substrate via a

facile double sided tape-assisted transfer approach using hollow Au-Ag alloy

nanourchins. Nanoscale 2014, 6 (5), 2567-72.

22. Jensen, T.; Kelly, L.; Lazarides, A.; Schatz, G. C., Electrodynamics of Noble

Metal Nanoparticles and Nanoparticle Clusters. J. Cluster Sci. 1999, 10 (2), 295-317.

23. Schatz, G. C., Electrodynamics of nonspherical noble metal nanoparticles and

nanoparticle aggregates. J. Mol. Struc.-THEOCHEM 2001, 573 (1-3), 73-80.

24. Etchegoin, P. G.; Le Ru, E. C., Basic Electromagnetic Theory of SERS. In

Surface Enhanced Raman Spectroscopy, Schlücker, S., Ed. Wiley-VCH: Weinheim,

Germany, 2011; pp 1-37.

25. Frens, G., Controlled Nucleation for the Regulation of the Particle Size in

Monodisperse Gold Suspensions. Nat. Phys. Sci. 1973, 241 (105), 20-22.

26. Leopold, N.; Lendl, B., A New Method for Fast Preparation of Highly Surface-

Enhanced Raman Scattering (SERS) Active Silver Colloids at Room Temperature by

Reduction of Silver Nitrate with Hydroxylamine Hydrochloride. J. Phys. Chem. B 2003,

107 (24), 5723-5727.

Page 56: Raman Spectroscopy Substrate Optimization and Method

43

27. Sharma, B.; Fernanda Cardinal, M.; Kleinman, S. L.; Greeneltch, N. G.;

Frontiera, R. R.; Blaber, M. G.; Schatz, G. C.; Van Duyne, R. P., High-performance

SERS substrates: Advances and challenges. MRS Bull. 2013, 38 (8), 615-624.

28. Sharma, B.; Cardinal, M. F.; Ross, M. B.; Zrimsek, A. B.; Bykov, S. V.;

Punihaole, D.; Asher, S. A.; Schatz, G. C.; Van Duyne, R. P., Aluminum Film-Over-

Nanosphere Substrates for Deep-UV Surface-Enhanced Resonance Raman Spectroscopy.

Nano Lett. 2016, 16 (12), 7968-7973.

29. Lin, W.-C.; Tsai, T.-R.; Huang, H.-L.; Shiau, C. Y.; Chiang, H.-P., SERS Study

of Histamine by Using Silver Film over Nanosphere Structure. Plasmonics 2012, 7 (4),

709-716.

30. DeJesus, J. F.; Trujillo, M. J.; Camden, J. P.; Jenkins, D. M., N-Heterocyclic

Carbenes as a Robust Platform for Surface-Enhanced Raman Spectroscopy. J. Am. Chem.

Soc. 2018, 140 (4), 1247-1250.

31. Cui, S.; Dai, Z.; Tian, Q.; Liu, J.; Xiao, X.; Jiang, C.; Wu, W.; Roy, V. A. L.,

Wetting properties and SERS applications of ZnO/Ag nanowire arrays patterned by a

screen printing method. J. Mater. Chem. C 2016, 4 (26), 6371-6379.

32. Huang, J.; Ma, D.; Chen, F.; Bai, M.; Xu, K.; Zhao, Y., Ag Nanoparticles

Decorated Cactus-Like Ag Dendrites/Si Nanoneedles as Highly Efficient 3D Surface-

Enhanced Raman Scattering Substrates toward Sensitive Sensing. Anal. Chem. 2015, 87,

10527-10534.

33. Liu, J.; Meng, G.; Li, Z.; Huang, Z.; Li, X., Ag-NP@Ge-nanotaper/Si-

micropillar ordered arrays as ultrasensitive and uniform surface enhanced Raman

scattering substrates. Nanoscale 2015, 7 (43), 18218-24.

Page 57: Raman Spectroscopy Substrate Optimization and Method

44

34. Xu, K.; Fu, C.; Gao, Z.; Wei, F.; Ying, Y.; Xu, C.; Fu, G., Nanomaterial-based

gas sensors: A review. Instrum. Sci. Technol. 2017, 46 (2), 115-145.

35. Agapov, R. L.; Srijanto, B.; Fowler, C.; Briggs, D.; Lavrik, N. V.; Sepaniak, M.

J., Lithography-free approach to highly efficient, scalable SERS substrates based on

disordered clusters of disc-on-pillar structures. Nanotechnology 2013, 24 (50), 505302.

36. Le Ru, E. C.; Blackie, E.; Meyer, M.; Etchegoin, P. G., Surface Enhanced

Raman Scattering Enhancement Factors: A Comprehensive Study. J. Phys. Chem. C

2007, 111, 13794-13803.

37. Pavel, I. E.; Alnajjar, K. S.; Monahan, J. L.; Stahler, A.; Hunter, N. E.;

Weaver, K. M.; Baker, J. D.; Meyerhoefer, A. J.; Dolson, D. A., Estimating the

Analytical and Surface Enhancement Factors in Surface-Enhanced Raman Scattering

(SERS): A Novel Physical Chemistry and Nanotechnology Laboratory Experiment. J.

Chem. Educ. 2012, 89 (2), 286-290.

38. Colthup, N. B.; Daly, L. H.; Wiberley, S. E., Introduction to infrared and Raman

spectroscopy. Academic Press: San Diego, Calif., 1998.

39. Angel, S. M.; Carrabba, M.; Cooney, T. F., The utilization of diode lasers for

Raman spectroscopy. Spectrochim. Acta A: 1995, 51 (11), 1779-1799.

40. Hutley, M. C., Diffraction gratings. Academic Press: London, 1982.

41. Loewen, E. G.; Popov, E., Diffraction gratings and applications. M. Dekker: New

York, 1997.

Page 58: Raman Spectroscopy Substrate Optimization and Method

45

Chapter 3

Direct Surface Enhanced Raman Spectroscopic Detection of Cortisol at

Physiological Concentrations

Page 59: Raman Spectroscopy Substrate Optimization and Method

46

A version of this chapter was originally published in Analytical Chemistry by T. Joshua

Moore and Bhavya Sharma.

Moore, T. J.; Sharma, B., Direct Surface Enhanced Raman Spectroscopic Detection of

Cortisol at Physiological Concentrations. Anal Chem 2020, 92 (2), 2052-2057.

This article describes the detection of cortisol in two different matrices (ethanol and a

human serum mimic) at the physiologically relevant concentrations using surface enhanced

Raman spectroscopy. Computational models were employed to identify the vibrational

modes present in the normal Raman spectrum, and to correlate the peaks arising in the

SERS spectrum to the calculated peaks. Multivariate techniques (i.e. principal component

analysis) were applied to samples in a complex medium. The article was edited by all

authors and was subjected to peer review.

Page 60: Raman Spectroscopy Substrate Optimization and Method

47

3.1 Abstract

Cortisol is an important steroid hormone in human physiology. Variations or

abnormalities in the physiological cortisol levels control acute and chronic stress response,

as well as contribute to diseases and syndromes including Addison’s disease and Cushing

syndrome. The ability to monitor cortisol levels in the physiological range is key in

diagnosis and monitoring of these conditions, where current methodology for

determination of cortisol levels relies on instrumentation that requires extensive sample

preparation, long run times, and is destructive to the sample. Raman spectroscopy provides

rapid sample analysis with relatively simple instrumentation; however, Raman

spectroscopy is an inherently weak technique. To provide an enhanced Raman signal, we

use surface enhanced Raman spectroscopy (SERS) which utilizes oscillating electric fields

of metal nanoparticles, enhancing the overall electric field and therefore resulting in an

enhanced signal. We demonstrate SERS-based detection of cortisol in the physiologically

relevant range using colloidal silver nanoparticles in ethanolic solutions and bovine serum

albumin. The SERS spectra obtained in an ethanol matrix demonstrate a sigmoidal

concentration response over the physiologically relevant concentration range, with a limit

of detection established at 177 nM. Analysis of cortisol solutions in a complex matrix

(bovine serum albumin in phosphate buffered saline) is also demonstrated through the use

of principal components analysis, a multivariate technique, which shows the separation of

cortisol in a linear fashion with respect to cortisol concentration.

Page 61: Raman Spectroscopy Substrate Optimization and Method

48

3.2 Introduction

The physiological stress response begins with the presentation of a stimulus

(stressor) that generates a stress perception response in the brain which activates the

appropriate systems in the body.1 This neuroendocrine response cascade2 proceeds through

a pair of pathways that take place along the hypothalamic-pituitary-adrenal (HPA) axis,

resulting in the formation of catecholamine neurotransmitters (dopamine, epinephrine, and

norepinephrine) and cortisol.3, 4 Cortisol, which is produced in the zona fasciculata of the

adrenal cortex, has long been recognized as the primary biomarker in evaluation of stress-

related disorders.5 Normally, cortisol exhibits a diurnal rhythm with serum levels showing

rapid elevation during the first 30-45 minutes after awakening (the cortisol awakening

response) followed by a decline over the course of the day.6, 7

Cortisol, a glucocorticoid steroid hormone derived from cholesterol, is the product

of a biological synthetic pathway that includes a series of enzyme-mediated

dehydrogenation and hydroxylation reactions. The function of cortisol in vertebrates is

multi-faceted; in addition to its role as a stress biomarker, cortisol plays a role in bone

metabolism,8 maintenance of blood pressure,9 the immune system,10 and

gluconeogenesis.11 Cortisol serum levels demonstrate variability between males and

females and with time of day, but they are generally in the nanomolar range, with morning

levels ranging from 138-690 nM and evening levels ranging from 55-386 nM.12

Serum cortisol levels are sensitive to acute and chronic stress conditions and have

been measured by a variety of techniques, including radio immunoassay (RIA), enzyme-

linked immunosorbent assay (ELISA), chemiluminescence immunoassay (CLIA),

capillary electrophoresis-based immunoassay (CE-IA), gas chromatography–mass

Page 62: Raman Spectroscopy Substrate Optimization and Method

49

spectrometry (GC-MS) and high performance liquid chromatography–tandem mass

spectrometry (HPLC-MS/MS).5, 7 Serum cortisol levels are commonly assessed to

differentiate between disease states that result from either the overproduction (Cushing’s

syndrome) or underproduction (Addison’s disease) of cortisol.13

While the above techniques can provide quantitative determination of cortisol

levels in body fluids, they require extensive sample preparation and analysis times, can be

destructive to the samples, and rely on large, immobilized instrumentation. Raman

spectroscopy can overcome these disadvantages, as samples can be rapidly analyzed in any

physical state without destruction of the sample, and instrumentation can be miniaturized

to handheld systems. Additionally, Raman spectroscopy provides a wealth of chemical

information attributed to its excellent specificity. However, Raman scattering is an

inherently weak inelastic light scattering process, with most molecules having scattering

cross-sections on the order of 10-30 to 10-25 cm2 per molecule.14

A common way to overcome this weak signal is through the use of surface-

enhanced Raman spectroscopy (SERS). First observed in 1974 by Fleischman et al., and

subsequently explained by Jeanmaire and Van Duyne in 1977, SERS involves capitalizing

on the inherent properties of plasmonic metals.15, 16 When a molecule of interest is located

in close proximity to a plasmonic metal nanostructure and illuminated by incident laser

radiation, the conduction band electrons in the metal nanoparticle start to oscillate. This

phenomenon is referred to as the localized surface plasmon resonance (LSPR) and results

in an increased electric field being present at the surface of the metal nanostructure. This

electric field amplifies the Raman scattered light as well as scattering cross-sections, which

in SERS can exhibit an increase of 106 – 1014 in magnitude.17-19

Page 63: Raman Spectroscopy Substrate Optimization and Method

50

Detection of cortisol using Raman-based methodology is an emergent field that has

so far been limited to quantitative resonance Raman methods20 and paper microdevices.21

These methods rely on indirect detection strategies, in which the measured Raman/SERS

signal is obtained from a reporter molecule instead of the analyte. Direct detection

strategies, however, are focused on the detection of the analyte itself. Herein, we report on

the direct SERS detection of cortisol in the physiologically relevant range.

3.3 Experimental Section

3.3.1 Materials and Reagents

Cortisol solid (≥98%) and phosphate buffered saline (PBS) powder were purchased

from Sigma-Aldrich. Silver nitrate solid (AgNO3, ≥99.7%), ethanol (EtOH, 99.5%), and

sodium hydroxide pellets (NaOH, >99.0%) were purchased from Fisher Scientific.

Hydroxylamine hydrochloride (HA, NH2OH · HCl, >99%) was purchased from ACROS

Organics. Bovine serum albumin was obtained from VWR All reagents were used as

received without further purification. All aqueous solutions were prepared using ultrapure

deionized water (DI-H2O, ≥ 18.0 MΩ cm).

3.3.2 Synthesis of Hydroxylamine-reduced Silver Nanoparticles

Hydroxylamine-reduced silver nanoparticles (HA-AgNPs) were prepared through

a modified version of synthetic methods described by Leopold and Lendl.22 In brief, 150

μL of 1 M HA, 300 μL of 1 M NaOH, and 89.550 mL of DI-H2O were combined, yielding

a reaction mixture with HA and NaOH concentrations of 1.67 mM and 3.33 mM,

respectively.23 The solution was placed on a magnetic stirrer (300 RPM) prior to the rapid

addition of a 10 mL volume of 10 mM AgNO3. The resulting colloids were characterized

Page 64: Raman Spectroscopy Substrate Optimization and Method

51

by scanning electron microscopy imaging (SEM, AURIGA series FIB/SEM, Zeiss) and the

collection of UV-visible extinction spectra (Cary 5000, Agilent). The resulting colloids

show an average diameter of ~30 nm, and the characteristic LSPR maximum was located

at 409 nm (Figures 3.1 and 3.2), concurring with literature values.22, 23 The enhancement

factor of the synthesized HA-AgNPs was calculated to be ~3 × 105.

3.3.3 SERS Sample Preparation

Cortisol is a sparingly soluble compound, with aqueous solubility limits on the

order of < 1 mM. To alleviate solubility concerns, an ethanolic solution (40 mM) of cortisol

was prepared, followed by serial dilution in EtOH, to achieve the desired physiologically

relevant concentrations. One mL aliquots of the synthesized HA-AgNPs were centrifuged

(6000 RPM, ~3200×g, 7 mins), the supernatant was discarded, and the resulting decanted

particles were combined with 250 μL of the cortisol solutions. The samples were incubated

for 2 hours before collection of SERS spectra. A blank, containing only the decanted HA-

AgNPs and EtOH, was also prepared and allowed to incubate for 2 hours to enable spectral

subtraction of the EtOH vibrational contributions.

Page 65: Raman Spectroscopy Substrate Optimization and Method

52

Figure 3.1. SEM image of the prepared hydroxylamine-reduced silver nanoparticles (HA-AgNP), showing good monodispersity and an average diameter of ~ 30 nm.

Page 66: Raman Spectroscopy Substrate Optimization and Method

53

Figure 3.2. Extinction spectrum of synthesized HA-AgNP, showing an LSPR maximum located at 409 nm.

Page 67: Raman Spectroscopy Substrate Optimization and Method

54

3.3.4 Normal Raman and SERS Spectral Acquisition

Normal Raman and SERS spectra were obtained using a home-built confocal

Raman instrument. The 532 nm excitation wavelength (λex) was provided by an optically-

pumped semiconductor laser (Coherent). The excitation light was directed through a series

of optics and opto-mechanics (Thorlabs) before entering the back of an inverted Nikon Ti-

u microscope equipped with a 40x objective (Nikon, NA 0.60). The 180° backscattered

Raman light was collected through the same objective prior to passing through a dichroic

mirror (Chroma Technology Corporation). The Raman light was then focused through a

holographic notch filter (Semrock) onto the 100 μm slit of an IsoPlane SCT-320

spectrometer (Princeton Instruments) equipped with a PIXIS 400 CCD camera (Princeton

Instruments). Spectral resolution for the spectrometer was 4 cm-1. The obtained spectra

were processed using GRAMS/AI software (ThermoScientific); spectral processing

included blank subtraction, baseline and offset correction, and Savitsky-Golay smoothing.

3.3.5 Theoretical Raman Calculations

Normal Raman spectra for cortisol were calculated using density functional theory

(DFT) with the Becke, three-parameter, Lee-Yang-Parr (B3LYP)24 hybrid functional and

the 6-31G* basis set. Cortisol geometries were optimized by using the basis set and the

optimized geometries were utilized to perform Raman polarizability calculations in

response to a perturbing electric field.25-29 The resulting Raman polarizabilities were used

to calculate the theoretical Raman spectra of the cortisol molecule. Calculations were

carried out using the NWCHEM computational chemistry package (Release 6.6).30 The

resulting Raman spectrum and normal modes were visualized with Chemcraft software.31

Page 68: Raman Spectroscopy Substrate Optimization and Method

55

Vibrational mode assignments were completed through analysis of the visualized

molecular motions for each normal mode.

3.4 Results and Discussion

The structure of cortisol (Figure 3.3a) and the cholesterol ring numbering

convention (Figure 3.3b), including the nomenclature for identifying the individual ring

moieties in the cholesterol skeleton, are shown in Figure 3.3.

A sample of solid cortisol was placed between two glass coverslips and placed on

the confocal Raman microscope. The normal Raman spectrum was collected from 200-

4000 cm-1 at 532 nm excitation, with a power of 10 mW and 10 s of acquisition time. The

fingerprint region from 200-1800 cm-1 is shown (Figure 3.4b). The computed Raman

spectrum (Figure 3.4a), which used B3LYP/6-31G* DFT, is compared to the experimental

spectrum in this region. The full spectrum (200 – 4000 cm-1) and a closer view of the high

wavenumber region (2800 – 3100 cm-1) can be found in Figures 3.5 and 3.6, respectively.

The calculated vibrational frequencies were scaled by a factor of 0.9613 for the

combination of exchange correlation functional and basis set,32 and the resulting

vibrational modes and intensities were Lorentzian broadened by 5 cm-1. Though the

intensities of calculated modes appear to deviate from what is observed experimentally, the

wavenumber position of the normal modes after scaling agree with the experimental

spectrum. The computed vibrational normal modes were visualized with Chemcraft

software to elucidate the atomic motions responsible for the vibrational frequencies. The

peak located at 1609 cm-1 is representative of a C=C stretching mode, while the peak at

1641 cm-1 corresponds to the C=O stretching in the A ring of the cortisol molecule.

Page 69: Raman Spectroscopy Substrate Optimization and Method

56

Figure 3.3. Structure of cortisol (a) and the carbon-numbering and ring identification conventions for the cholesterol backbone present in cortisol (b).

Page 70: Raman Spectroscopy Substrate Optimization and Method

57

Figure 3.4. Theoretical (a) and experimental (b) normal Raman spectra of cortisol solid in the region from 200−1800 cm−1. There is good agreement between the theoretical and experimental spectra. For the experimental spectrum, λex = 532 nm, P = 10 mW, and t = 10 s.

Page 71: Raman Spectroscopy Substrate Optimization and Method

58

Figure 3.5. Experimental (top, red) and theoretical (bottom, blue) normal Raman spectra for cortisol in the region from 200 – 4000 cm-1. For the experimental spectrum, λex = 532 nm, P = 10 mW, and t = 10 s.

Page 72: Raman Spectroscopy Substrate Optimization and Method

59

Figure 3.6. Theoretical (a) and experimental (b) normal Raman spectra for cortisol in the high wavenumber region.

Page 73: Raman Spectroscopy Substrate Optimization and Method

60

The region from 600-950 cm-1 consists of normal modes related to the vibration of the

cholesterol skeleton moiety. The observed vibrational modes from the experimental

spectrum, the corresponding theoretical vibrational modes, and mode assignments are

summarized in Table 3.1

SERS spectra of cortisol were then collected at concentrations ranging from 125

nM to 1000 nM. Briefly, three exposures of 30 s each were collected for each concentration

of cortisol and for the HA-AgNP/EtOH blank after incubation. The frame-averaged blank

spectra were subtracted from the frame-averaged SERS spectra of cortisol and the resulting

spectra are shown in Figure 3.7. The pre-dominant SERS bands observed in these samples

occur at 1146, 1269, 1344, 1378, 1432, 1450, and 1504 cm-1 (Table 3.1). A less intense

band is also observed at 1073 cm-1, corresponding to an asymmetric ring breathing mode

of the C ring of cortisol. In SERS measurements, it is common to see peaks present in the

normal Raman spectrum shifted by an amount from 1-30 cm-1.33, 34 The 1146 cm-1 band

corresponds to the CH2 twist + CH bend + OH bend modes located at 1149

cm-1 in the normal Raman spectrum. The SERS band observed at 1269 cm-1 is well

enhanced from the solid spectrum, where it is predicted at 1270 cm-1 but with low intensity.

This peak is the result of a combination of CH bend + CH2 wag + OH bend modes.

Page 74: Raman Spectroscopy Substrate Optimization and Method

61

Table 3.1. Theoretical normal Raman, experimental normal Raman, and SERS peak positions and vibrational modes for cortisol.

Theoretical, scaled (cm-1)

Normal Raman, solid

(cm-1)

SERS (cm-1)

Vibrational Modes

1078 1077 1073 Asymmetric ring breathing (C ring) 1148 1149 1146 CH2 twist + CH bend + OH bend 1221 1222

1224 - C-C-C asymmetric stretch (D ring) C-C-C asymmetric stretch (A ring)

1242 1239 - C-C-C asymmetric stretch (C ring) 1270 1275 1269 CH bend + CH2 bend + OH bend 1332 1330 - C-C-C asymmetric stretch A ring 1341 1342

1346 1344 C-C-C symmetric stretch (A ring)

1381 1385

1382 1390

1378 CH3 symmetric bend + CH wag + OH wag (C ring)

1435 1432 1432 CH2 scissor (A, B rings) 1446 1448 1450 CH2 scissor (D ring + pendant) 1611 1609 - C=C stretch

- - 1504 C=C stretch 35 OR C-C-C asymmetric ring deform.36

1704 1641 - C=O stretch

Page 75: Raman Spectroscopy Substrate Optimization and Method

62

Figure 3.7. SERS spectra (1000 – 1600 cm-1) of nM cortisol solutions representative of the physiological normal range in serum. Spectra were collected with λex = 532 nm, P = 10 mW, t = 30 s.

Page 76: Raman Spectroscopy Substrate Optimization and Method

63

The band located at 1344 cm-1 arises from the C-C-C symmetric stretching in the

A ring. The less intense band located at 1378 cm-1 arises from a combination of two

theoretical bands (1381 and 1385 cm-1, scaled) resulting from CH3 symmetric bending

motions in opposite directions, coupled with CH bend and OH bend vibrations. These two

vibrational modes convolve into one band located in the solid spectrum at 1382 cm-1.

The SERS peak located at 1432 cm-1 (CH2 scissoring on the A and B rings) matches

the peak position observed in the solid spectrum, while a slight upshift in frequency is

observed for the SERS band located at 1450 cm-1 (1448 cm-1 in the solid, CH2 scissoring

of the D ring at the attached pendant).

The predominant peak of the SERS spectrum is located at 1504 cm-1, which is

neither an observed peak in the experimental solid spectrum or a theoretical peak predicted

by DFT. The nearest observed normal Raman peaks are the peaks located at 1609 cm-1

(C=C stretch) and 1641 cm-1 (C=O stretch). If the cortisol molecule is oriented onto the Ag

surface such that the lone pairs on the double-bonded oxygen on the A ring are donated

into the silver d-orbitals, the C=O stretch vibration would be significantly damped to the

point of muting the vibration. Due to the large magnitude of the frequency downshift (-137

cm-1) that would be required to assign the 1504 cm-1 vibration to the C=O stretch, it seems

unlikely that the peak arises from this vibration. Methylene blue (MB, Figure 3.8) is an

azo dye with a fused 6-member ring system. This molecule also shows a SERS band located

at 1504 cm-1, which Quester et al. identify as an asymmetric C-C-C ring deformation.37

Since the only locations in MB that have these C-C-C regions are the peripheral rings, each

of which contain C=C and C-C-C functionalities, we find it reasonable that the observed

Page 77: Raman Spectroscopy Substrate Optimization and Method

64

Figure 3.8. Structure of methylene blue.

Page 78: Raman Spectroscopy Substrate Optimization and Method

65

SERS band could be attributed to either a significant downshift of the C=C bond vibration,

or to a C-C-C deformation in the cholesterol skeleton.

The integrated areas were plotted and fit with a Boltzmann sigmoidal function,

resulting in adjusted R2 values of 0.98613 for the 1269 cm-1 peak, and 0.99269 for the 1504

cm-1 peak (Figure 3.9). These peaks were selected for analysis due to their increased SERS

intensity and the absence of overlap with other SERS bands. The adjusted R2 values

demonstrate that the sigmoidal fit obtained from the Boltzmann function agree very well

with the obtained signal. The other adjusted R2 values for the peaks not shown in Figure

3.9 are 0.99628 (1073 cm-1), 0.94496 (1146 cm-1), 0.98613 (1344 cm-1), 0.99721 (1432 cm-

1), and 0.99397 (1450 cm-1), demonstrating a sigmoidal concentration dependent SERS

response for these peaks, which agrees with precedence in the scientific literature for SERS

and plasmonic biosensors. 38-41 The visual limit of detection (LOD) can be estimated from

the plots in Figure 3.9 to be in the range from 125 – 250 nM. To determine the limit of

detection (LOD) for cortisol, the 1504 cm-1 peak was selected as the peak of interest since

has an R2 value closer to unity in the linear regression analysis. Three frames at each of the

concentrations were collected and the peak area was integrated for each spectrum. The

resulting concentration and integrated area data were analyzed using ANOVA, and the

standard error of the intercept was calculated. The LOD (3.3 x standard error/slope) was

calculated to be 177 nM, which agrees with visual LOD analysis of ~ 200 nM.

The linearity of the concentration response of the SERS spectra was investigated

by peak fitting the individual 30 s exposures for each concentration of cortisol. The peaks

were fit using GRAMS/AI software (ThermoScientific), and the resultant area values were

Page 79: Raman Spectroscopy Substrate Optimization and Method

66

Figure 3.9. Sigmoidal concentration response plots obtained after peak fitting triplicate SERS spectra of physiological cortisol concentrations for the 1269 cm-1 (top) and 1504 cm-1 (bottom) peaks.

Page 80: Raman Spectroscopy Substrate Optimization and Method

67

analyzed statistically. Each peak’s mean area was plotted (Origin software) as a function

of concentration, and error bars were established as ± 1 standard deviation from the mean.

The ensemble of sigmoidal response and linearity plots are shown in Figures 3.10 and 3.11.

Error bars are present on all points in Figures 3.9-11, but some are so small in magnitude

that they are obscured by the data marker.

Though promising, the above results were obtained in ethanolic solutions, which is

a vastly different solvent environment than found in vivo. Biological samples are more

complex, with serum composed of a mixture of proteins, salts and other small- and

macromolecule interferents. This results in SERS spectra of analyte molecules in complex

media that are a combination of the SERS spectrum of the analyte molecule with the SERS

spectrum of the complex medium. Often, the SERS signals of the interferents from the

media overwhelm the SERS peaks that arise from the analyte molecule. We tested the

SERS-based detection scheme developed for ethanolic cortisol solutions in complex media

to determine the applicability of this sensing platform to a more representative matrix for

human samples. As our complex medium, we utilized a solution of bovine serum albumin

(BSA), a globular protein found in blood plasma which is chemically similar to human

serum albumin, in phosphate buffered saline (PBS). One mL aliquots of the HAAgNP were

centrifuged and decanted as described above, and then were incubated for 2 hours in

mixtures containing variable cortisol concentrations (2 μM, 1 μM, 750 nM, 500 nM, 250

nM, and 125 nM) with fixed BSA-PBS concentrations (0.75% BSA by mass). Since

biological samples are more complex than those produced in vitro, we sought to mimic this

complexity by detecting cortisol in a matrix with interferents that would be more

Page 81: Raman Spectroscopy Substrate Optimization and Method

68

Figure 3.10. Sigmoidal concentration response plots for the observed peaks in SERS spectra of cortisol obtained from samples in the physiological range (125-1000 nM). Error bars are present on all data points, but the small magnitude of some error bars is obscured by the data marker in several instances.

Page 82: Raman Spectroscopy Substrate Optimization and Method

69

Figure 3.11. Linear regression plots for the observed peaks in SERS spectra of cortisol

obtained from samples in the physiological range (125-1000 nM).

Page 83: Raman Spectroscopy Substrate Optimization and Method

70

representative of those found in a biological serum sample. Spectra were collected with

532 nm excitation, 10 mW power, and 60 s collection time. In this more complex matrix,

the cortisol signal was indeed suppressed due to the biological interferents, making a linear

regression analysis insufficient for these samples. Higher-order multivariate techniques are

able to discern differences in spectra that are not visible to the naked eye, and as such the

resulting spectra were analyzed using a custom MATLAB script for principal component

analysis (PCA), which generated a series of loadings for the principal components (PC).

The loading for PC1, shown in Figure 3.12b, was dominated by the SERS spectrum of

BSA.42 The loading for PC2 (Figure 3.12c), however, shows peaks that align well with

expected SERS bands of cortisol, especially the peaks at 1270, 1374, 1440, and 1517 cm-

1, which are only slightly shifted from their SERS positions in the ethanolic solution. The

difference in the solvent environment (aqueous vs. ethanolic), coupled with the presence

of a more complex, proteinaceous, and saline environment accounts for these shifts. The

PCA shows that the solutions containing cortisol clearly separate from the blank, which is

composed of just the BSA-PBS matrix (Figure 3.12a), indicating that these concentrations

can be detected with SERS, and that the signal from the samples is significantly different

than that of the BSA-PBS matrix. Furthermore, over the specified concentration range (125

nM – 2 μM), the samples separate in a linear fashion according to cortisol concentration

(PC2).

Page 84: Raman Spectroscopy Substrate Optimization and Method

71

Figure 3.12. Principal components analysis of cortisol in the physiologically relevant concentration range (low μM to 125 nM) in 0.75% bovine serum albumin-phosphate buffered saline (BSA-PBS). (a) Plot of scores for PC1 vs. PC2. The data point labels indicate the concentration of cortisol present in the sample. (b) The loading for PC1, which shows that PC1 is dominated by contributions from the BSA. (c) The loading for PC2, which shows peaks arising from cortisol.

Page 85: Raman Spectroscopy Substrate Optimization and Method

72

3.5 Conclusions

In conclusion, we have used DFT calculations on optimized molecular geometries

to generate the theoretical normal Raman spectrum of cortisol. Correlating the calculations

with the normal Raman spectrum of solid cortisol, we have assigned the normal modes for

the molecule through visualization of the resultant vibrations. Additionally, SERS spectra

for ethanolic cortisol solutions have been obtained over the physiologically relevant range

(125–1000 nM), and the spectra obtained from these SERS measurements show a linear

response. The LOD for ethanolic solutions was determined to be 177 nM. We are, to our

knowledge, the first to have directly detected cortisol via SERS without the use of a reporter

molecule. We have also demonstrated that this method can be used to detect cortisol in a

serum mimic at physiologically relevant concentrations through use of multivariate

analysis (PCA).

3.6 Acknowledgements

The authors would like to acknowledge the University of Tennessee, Knoxville’s

financial support through start-up funding. Computational work was carried out through

use of the Advanced Computing Facility at the University of Tennessee, Knoxville. The

authors would like to thank Dr. Lonnie Crosby for valuable conversations regarding the

use of NWChem software. The authors would like to acknowledge the University of

Tennessee JIAM Microscopy Center for instrument use, scientific, and technical

assistance.

Page 86: Raman Spectroscopy Substrate Optimization and Method

73

3.7 References

1. Dhabhar, F. S.; McEwen, B. S., Acute stress enhances while chronic stress

suppresses cell-mediated immunity in vivo: a potential role for leukocyte trafficking.

Brain Behav. Immun. 1997, 11 (4), 286-306.

2. Wolf, O. T.; Atsak, P.; de Quervain, D. J.; Roozendaal, B.; Wingenfeld, K.,

Stress and Memory: A Selective Review on Recent Developments in the Understanding

of Stress Hormone Effects on Memory and Their Clinical Relevance. J. Neuroendocrinol.

2016, 28 (8).

3. Eisenhofer, G.; Kopin, I. J.; Goldstein, D. S., Catecholamine metabolism: a

contemporary view with implications for physiology and medicine. Pharmacol. Rev.

2004, 56 (3), 331-49.

4. Randall, M., The Physiology of Stress: Cortisol and the Hypothalamic-Pituitary-

Adrenal Axis. Dartmouth Undergrad. J. Sci. 2011.

5. Jia, M.; Chew, W. M.; Feinstein, Y.; Skeath, P.; Sternberg, E. M.,

Quantification of cortisol in human eccrine sweat by liquid chromatography-tandem mass

spectrometry. Analyst 2016, 141 (6), 2053-60.

6. Clow, A.; Hucklebridge, F.; Thorn, L., The cortisol awakening response in

context. Int. Rev. Neurobiol. 2010, 93, 153-75.

7. Gaffey, A. E.; Bergeman, C. S.; Clark, L. A.; Wirth, M. M., Aging and the HPA

axis: Stress and resilience in older adults. Neurosci. Biobehav. Rev. 2016, 68, 928-945.

Page 87: Raman Spectroscopy Substrate Optimization and Method

74

8. Chiodini, I.; Torlontano, M.; Carnevale, V.; Trischitta, V.; Scillitani, A.,

Skeletal involvement in adult patients with endogenous hypercortisolism. J. Endocrinol.

Invest. 2014, 31 (3), 267-276.

9. Forhead, A. J.; Pipkin, F. B.; Fowden, A. L., Effect of cortisol on blood pressure

and the renin-angiotensin system in fetal sheep during late gestation. J. Physiol. 2000,

526 (1), 167-176.

10. Lange, T.; Dimitrov, S.; Born, J., Effects of sleep and circadian rhythm on the

human immune system. Ann. N. Y. Acad. Sci. 2010, 1193 (1), 48-59.

11. Khani, S.; Tayek, J. A., Cortisol increases gluconeogenesis in humans: its role in

the metabolic syndrome. Clin. Sci. 2001, 101 (6), 739-47.

12. Mayo Clinic Laboratories. Test ID: CINP Cortisol, Serum, LC-MS/MS.

https://www.mayocliniclabs.com/test-catalog/Clinical+and+Interpretive/9369 (accessed

January 13, 2019).

13. Turpeinen, U.; Hämäläinen, E., Determination of cortisol in serum, saliva and

urine. Best Pract. Res. Clin. Endocrinol. Metab. 2013, 27 (6), 795-801.

14. Kneipp, K.; Kneipp, H.; Itzkan, I.; Dasari, R. R.; Feld, M. S., Ultrasensitive

Chemical Analysis by Raman Spectroscopy. Chem. Rev. 1999, (99), 2957-2975.

15. Fleischmann, M.; Hendra, P. J.; McQuillan, A. J., Raman spectra of pyridine

adsorbed at a silver electrode. Chem. Phys. Lett. 1974, 26 (2), 163-166.

16. Jeanmaire, D. L.; Van Duyne, R. P., Surface Raman Spectroelectrochemistry Part

I. Heterocyclic, Aromatic, and Aliphatic Amines Adsorbed on the Anodized Silver

Electrode. J. Electroanal. Chem. 1977, 8, 1-20.

Page 88: Raman Spectroscopy Substrate Optimization and Method

75

17. Camden, J. P.; Dieringer, J. A.; Wang, Y.; Masiello, D. J.; Marks, L. D.;

Schatz, G. C.; Van Duyne, R. P., Probing the structure of single-molecule surface-

enhanced Raman scattering hot spots. J. Am. Chem. Soc. 2008, 130 (38), 12616-7.

18. Kneipp, K.; Kneipp, H.; Itzkan, I.; Dasari, R. R.; Feld, M. S., Surface-Enhanced

Raman Scattering and Biophysics. J. Phys.: Condens. Matter 2002, 14, R597-R624.

19. Sharma, B.; Frontiera, R. R.; Henry, A.-I.; Ringe, E.; Van Duyne, R. P., SERS:

Materials, applications, and the future. Mater. Today 2012, 15 (1-2), 16-25.

20. Gracie, K.; Pang, S.; Jones, G. M.; Faulds, K.; Braybrook, J.; Graham, D.,

Detection of cortisol in serum using quantitative resonance Raman spectroscopy. Anal.

Methods 2017, 9 (10), 1589-1594.

21. Gao Zhigang, Z. T., Deng Jiu, Li Xiaorui, Qu Yueyang, Lu Yao, Liu Tingjiao,

Luo Yong, Zhao Weijie, Lin Bingcheng, Ultrasensitive Detection of Hair Cortisol Based

on Portable Raman Spectrometer and Double-layer Paper Microdevice. Acta Chim.

Sinica 2017, 75 (4), 355-359.

22. Leopold, N.; Lendl, B., A New Method for Fast Preparation of Highly Surface-

Enhanced Raman Scattering (SERS) Active Silver Colloids at Room Temperature by

Reduction of Silver Nitrate with Hydroxylamine Hydrochloride. J. Phys. Chem. B 2003,

107 (24), 5723-5727.

23. Larmour, I. A.; Faulds, K.; Graham, D., SERS activity and stability of the most

frequently used silver colloids. J. Raman Spectrosc. 2012, 43 (2), 202-206.

24. Becke, A. D., Density‐functional thermochemistry. III. The role of exact

exchange. J. Chem. Phys. 1993, 98 (7), 5648-5652.

Page 89: Raman Spectroscopy Substrate Optimization and Method

76

25. Autschbach, J., Computation of Optical Rotation using Time–Dependent Density

Functional Theory. Computing Letters 2007, 3 (2), 131-150.

26. Autschbach, J., Time-dependent density functional theory for calculating origin-

independent optical rotation and rotatory strength tensors. ChemPhysChem 2011, 12 (17),

3224-35.

27. Jensen, L.; Autschbach, J.; Schatz, G. C., Finite lifetime effects on the

polarizability within time-dependent density-functional theory. J. Chem. Phys. 2005, 122

(22), 224115.

28. Krykunov, M.; Autschbach, J., Calculation of optical rotation with time-periodic

magnetic-field-dependent basis functions in approximate time-dependent density-

functional theory. J. Chem. Phys. 2005, 123 (11), 114103.

29. Krykunov, M.; Banerjee, A.; Ziegler, T.; Autschbach, J., Calculation of Verdet

constants with time-dependent density functional theory: implementation and results for

small molecules. J. Chem. Phys. 2005, 122 (7), 074105.

30. Valiev, M.; Bylaska, E. J.; Govind, N.; Kowalski, K.; Straatsma, T. P.; Van

Dam, H. J. J.; Wang, D.; Nieplocha, J.; Apra, E.; Windus, T. L.; de Jong, W. A.,

NWChem: A comprehensive and scalable open-source solution for large scale molecular

simulations. Comput. Phys. Commun. 2010, 181 (9), 1477-1489.

31. Zhurko, G. Chemcraft - graphical program for visualization of quantum

chemistry computations, 1.8 (build 536a); Ivanovo, Russia.

32. Merrick, J. P.; Moran, D.; Radom, L., An evaluation of harmonic vibrational

frequency scale factors. J. Phys. Chem. A 2007, 111 (45), 11683-700.

Page 90: Raman Spectroscopy Substrate Optimization and Method

77

33. Sharma, B.; Cardinal, M. F.; Ross, M. B.; Zrimsek, A. B.; Bykov, S. V.;

Punihaole, D.; Asher, S. A.; Schatz, G. C.; Van Duyne, R. P., Aluminum Film-Over-

Nanosphere Substrates for Deep-UV Surface-Enhanced Resonance Raman Spectroscopy.

Nano Lett 2016, 16 (12), 7968-7973.

34. Stacy, A. A.; Van Duyne, R. P., Surface enhanced raman and resonance raman

spectroscopy in a non-aqueous electrochemical environment: Tris(2,2′-

bipyridine)ruthenium(II) adsorbed on silver from acetonitrile. Chem. Phys. Lett. 1983,

102 (4), 365-370.

35. Patterson, M. L.; Weaver, M. J., Surface-enhanced Raman spectroscopy as a

probe of adsorbate-surface bonding: simple alkenes and alkynes adsorbed at gold

electrodes. J. Phys. Chem. 1985, 89 (23), 5046-5051.

36. Quester, K.; Avalos-Borja, M.; Vilchis-Nestor, A. R.; Camacho-López, M. A.;

Castro-Longoria, E., SERS Properties of Different Sized and Shaped Gold Nanoparticles

Biosynthesized under Different Environmental Conditions by Neurospora crassa Extract.

PLoS One 2013, 8 (10).

37. Quester, K.; Avalos-Borja, M.; Vilchis-Nestor, A. R.; Camacho-Lopez, M. A.;

Castro-Longoria, E., SERS properties of different sized and shaped gold nanoparticles

biosynthesized under different environmental conditions by Neurospora crassa extract.

PLoS One 2013, 8 (10), e77486.

38. Altun, A. O.; Bond, T.; Pronk, W.; Park, H. G., Sensitive Detection of

Competitive Molecular Adsorption by Surface-Enhanced Raman Spectroscopy.

Langmuir 2017, 33 (28), 6999-7006.

Page 91: Raman Spectroscopy Substrate Optimization and Method

78

39. Chowdhury, M. H.; Gant, V. A.; Trache, A.; Baldwin, A.; Meininger, G. A.;

Coté, G. L., Use of surface-enhanced Raman spectroscopy for the detection of human

integrins. J. Biomed. Opt. 2006, 11 (2).

40. Haes, A. J.; Van Duyne, R. P., A nanoscale optical biosensor: sensitivity and

selectivity of an approach based on the localized surface plasmon resonance spectroscopy

of triangular silver nanoparticles. J. Am. Chem. Soc. 2002, 124 (35), 10596-604.

41. Muehlig, A.; Jahn, I. J.; Heidler, J.; Jahn, M.; Weber, K.; Sheen, P.; Zimic,

M.; Cialla-May, D.; Popp, J., Molecular Specific and Sensitive Detection of

Pyrazinamide and Its Metabolite Pyrazinoic Acid by Means of Surface Enhanced Raman

Spectroscopy Employing In Situ Prepared Colloids. Appl. Sci. 2019, 9 (12).

42. Fazio, B.; D’Andrea, C.; Foti, A.; Messina, E.; Irrera, A.; Donato, M. G.;

Villari, V.; Micali, N.; Maragò, O. M.; Gucciardi, P. G., SERS detection of

Biomolecules at Physiological pH via aggregation of Gold Nanorods mediated by Optical

Forces and Plasmonic Heating. Sci. Rep. 2016, 6 (1).

Page 92: Raman Spectroscopy Substrate Optimization and Method

79

Chapter 4

Surface Enhanced Raman-based Detection of Volatile Organic

Compounds Using Porous Silicon Oxide Coated Disc-on-Pillar Arrays

Page 93: Raman Spectroscopy Substrate Optimization and Method

80

Portions of this chapter were originally published as a conference paper for The Optical

Society of America by Terence J. Moore and Bhavya Sharma.

Moore, T. J.; Sharma, B., Volatile Organic Compound Detection using Porous-Silicon-

Oxide Coated Disc-on-Pillar Arrays. In Advanced Photonics 2018 (BGPP, IPR, NP,

NOMA, Sensors, Networks, SPPCom, SOF), 2018.

This chapter describes the nanofabrication of a 3-dimensional SERS substrate, a porous

silicon oxide coated disc-on-pillar array, for the detection of volatile organic compounds

in the gas phase. Substrate optimization steps were performed to generate the best

enhancing substrate. The substrate was used to detect four VOC compounds in the gas

phase. Two of the compounds, benzenethiol and pyridine, readily adsorb to the surface of

the substrate and can be detected with miniaml exposure time. The gold disc-on-pillar array

demonstrates parts-per-million detection for benzenethiol vapor. For non-adsorbing

analytes, the disc-on-pillar array requires longer contact time, but still demonstrates

detection of the analytes. The substrate shows promise as an on-line monitoring method

for gas phase organic compounds.

This chapter is in preparation for submission to Analytical Chemistry in Spring 2020.

Page 94: Raman Spectroscopy Substrate Optimization and Method

81

4.1 Abstract

Volatile organic compounds (VOCs) are a large class of molecules that have low

boiling points and are capable of evaporation at room temperature. Exposure to VOCs can

have damaging impacts on human health and can give rise to environmental issues.

Detection of these molecules is important for a wide variety of industries; however typical

detection methods require large, non-portable instrumentation with long run times.

Surface-enhanced Raman spectroscopy (SERS) provides an attractive alternative to

conventional methods because it is rapid, sensitive, has high chemical specificity, and can

be combined with portable instrumentation. Herein we report on the characterization and

application of a SERS substrate consisting of plasmonic gold discs situated on top of pillars

(Au-DOP) for the detection of off-gassed VOCs at room temperature. We demonstrate the

detection of four molecules (benzenethiol, pyridine, benzene, and ortho-xylene) by the

substrate, with benzenethiol showing parts-per-billion (ppb) detection limits.

4.2 Introduction

Volatile organic compounds (VOCs) are a class of molecules that exhibit high

vapor pressure at normal room temperatures, resulting in low boiling points that allow a

large fraction of molecules to leave the bulk through evaporation. VOCs have been shown

to have adverse impact on human health, with health effects from exposure including

irritation of the nasopharyngeal tissues and eyes, nausea, loss of coordination, headaches,

and damage to the liver, kidney, and central nervous systems.1 Additionally, some VOCs

have been identified as either known or likely carcinogens, or have been shown to have

tumor-promoting effects.2 VOCs can also be produced during disease states, and as such

Page 95: Raman Spectroscopy Substrate Optimization and Method

82

their detection in the gas phase as biomarkers shows promise as a technique in medical

diagnostics.3

Indoor air quality monitoring strategies must include VOC detection due to a wide

range of sources of these compounds, including those that evaporate from cleaning

supplies, furniture, pesticides, paints, and varnishes.4 The major sources of VOCs include

petroleum refineries, chemical industries, automobile industries, textile manufacturing, and

solvent processes, which contribute to air pollution on both local and global scales.5 When

released into ambient air, VOCs become a primary contributor to photochemical smog by

reaction with nitrogen oxides (NOx) to form ozone and secondary organic aerosols.5-7 Thus,

the ability to monitor VOC emissions in real time is desirable for industrial and

environmental applications.

VOC detection is currently achieved through high-performance liquid

chromatography (HPLC) 8, 9 and gas chromatography-mass spectrometry (GC-MS);3, 10, 11

however, these techniques can be destructive, require long analysis times, lack chemical

specificity, and require expensive instrumentation. Nanomaterial-based gas sensors offer a

low-cost option that does not sacrifice sensitivity, with increased surface area to volume

ratios and porous structures.12 For these reasons, there has been increased recent interest in

development of nanomaterial based VOC sensors.3, 12-15

Raman spectroscopy provides an attractive alternative to the above methods

because it is rapid, specific, requires minimal or no sample preparation, is nondestructive

to the sample, and can be measured with portable instrumentation. However, normal

Raman spectroscopy suffers from being an inherently weak technique due to small Raman

scattering cross-sections for most molecules, which makes the technique less robust for

Page 96: Raman Spectroscopy Substrate Optimization and Method

83

ultrasensitive analyses.16 Since the permissible exposure limits for many VOCs are on the

order of parts-per-million,17 it is necessary to have sensitive analytic techniques in place to

monitor these compounds in real-time.

The weakness of normal Raman signal can be mitigated through taking advantage

of the localized surface plasmon resonance (LSPR), a phenomenon exhibited by noble

metal nanostructures (most commonly Au and Ag). When these metals are placed into the

incident electric field of a laser, the resultant oscillating conduction band electrons produce

an increased electric field surrounding the metal nanoparticle. Placement of the target

molecule within 1-2 nm18 -of a plasmonic metal surface or nanostructure takes advantage

of this increased electric field (the “hot spot”), resulting in an amplified Raman signal in a

technique known as surface-enhanced Raman spectroscopy (SERS). Theoretical

calculations for this electromagnetic enhancement reach ~ 1011, which allows for trace

analysis.19, 20

SERS substrates have been employed for detection of gas phase VOCs, including

explosives,21, 22, biomarkers of lung cancer,23 and BTEX chemicals (benzene, toluene,

ethylbenzene, and the xylenes).24 Gas detection schema include using substrates at cooled

temperatures which allows for condensation of the vapor onto the substrate.25, 26 In cases

where spectral acquisition is carried out at room temperature, higher powers and longer

acquisition times are typically used.27, 28 Herein, we report on the optimization of a SERS-

based substrate consisting of gold disc-on-pillar (Au-DOP) arrays that are coated with a

porous silicon oxide (PSO) layer to concentrate gas molecules into the SERS hot spots,

which permit rapid detection times when exposed to vapors from adsorbing analyte

molecules.

Page 97: Raman Spectroscopy Substrate Optimization and Method

84

4.3 Experimental Methods

4.3.1 Reagents

Benzene (≥99.0%) was obtained from Sigma-Aldrich. Benzenethiol (BT, 99%),

isopropanol (IPA, 99.6%), ethanol (EtOH, 99.5%), and pyridine (99+%) were obtained

from ACROS Organics. Ortho-xylene (99%) was obtained from Alfa Aesar. All chemicals

were used as received, without further purification steps. Aqueous solutions were prepared

using ultra-pure deionized water (>18 MΩ cm).

4.3.2 Nanofabrication

Disc-on-pillar (DOP) arrays were fabricated according to previously established

methods.29 Briefly, a vacuum evaporator equipped with dual electron beams (Thermonics

Laboratory, VE-240) was used to deposit a 5 nm Pt film via physical vapor deposition

(PVD) onto a (100) Si wafer with 100 nm layer of thermally-grown SiO2 on the surface.

After film deposition, the wafers were rapidly heated to ~850 °C over an 8 s period using

a furnace equipped with a radiative heat source set to maximum power (22 kW, First Nano,

Easy Tube 3000). The atmosphere inside the furnace was composed of a 10:1 mixture of

argon and hydrogen, respectively, and was maintained at a pressure of 735 torr. The thermal

processing step resulted in the dewetting of the Pt film and generated a stochastic array of

circular metal islands. These islands were used as masks for a two-step reactive ion etching

(RIE) process, both of which were completed using an Oxford PlasmaLab (Oxford

Instruments, UK) system with combination inductively- and capacitively-coupled plasma

(ICP/CCP). The first RIE step involved etching the SiO2 layer in a mixture of C4F8 and O2

(45 sccm and 2 sccm, respectively) gases at 7 mTorr pressure for 60 s. The subsequent RIE

Page 98: Raman Spectroscopy Substrate Optimization and Method

85

step focused on the etching of the Si in a mixture of SF6:C4F8:Ar at 10 mTorr for 60 s, with

flow rates of 56, 25, and 5 sccm, respectively. To reveal the stochastic array of bare Si

pillars, the wafers were immersed for 2 minutes in a 6:1 NH4F:HF buffered oxide etchant

(BOE) bath, resulting in the removal of any remaining SiO2 and the Pt RIE masks.

Immediately after the BOE wet etch, a 20 nm layer of SiO2 was deposited via atomic layer

deposition (ALD) using an Oxford FlexAL (Oxford Instruments, Oxford, UK), followed

by deposition of a 25 nm Au layer via PVD. A 2 nm layer of alumina (Al2O3) was deposited

on the wafers via PVD, followed by low temperature (~27 °C) plasma-enhanced chemical

vapor deposition (PECVD, Plasmalab System 100, Oxford Instruments) of a porous silicon

dioxide (PSO) layer in a mixture of silane (5% SiH4, 75 sccm) and nitrous oxide (N2O, 600

sccm) at 600 mTorr. Substrates were diced into 5 × 5 mm chips using an Accretech dicing

saw equipped with a silicon blade.

4.3.3 Off-gassing Experiments and SERS Spectra Collection

Off-gassing experiments were completed by suspending the fabricated substrates

in the headspace above a sample of neat analyte. After contact with analyte vapors for the

specified time, the substrates were removed from their vessels and transferred to a

microscope cover slide for SERS analysis. SERS measurements were obtained using a

home-built confocal Raman system in a 180° backscattering geometry. Excitation light

from a 785 nm diode laser (Innovative Photonic Solutions, Monmouth, NJ) was directed

through a series of optics into the back of an inverted Nikon Ti-U microscope and was then

focused through a 40× objective (Nikon, NA 0.60) onto the substrate surface. The Raman

scattered light was collected via the objective and focused onto the 100 μm slit of an Iso-

Page 99: Raman Spectroscopy Substrate Optimization and Method

86

Plane 320SCT spectrometer equipped with a PIXIS 400 charge-coupled-device (CCD)

camera (6 cm-1 resolution, both from Princeton Instruments). Spectra were acquired using

LightField software from Princeton Instruments, and the raw spectra were processed

(baseline correction, cosmic ray removal, background subtraction, and peak fitting) using

GRAMS/AI 9.3 software from ThermoScientific.

4.4 Results and Discussion

Previously, Agapov et al. reported on the fabrication of the Au-DOP arrays, which

demonstrate SERS enhancement factors on the order of 107.29 The surface architecture of

this Au-DOP arrays was investigated to determine if the presence of the pillars was

responsible for the SERS enhancement, or if the enhanced signal was just a by-product of

the deposition of a plasmonic metal surface. SERS enhancement is increased on roughened

surfaces due to the generation of hot spots from the surface features, and roughening of

substrate surfaces has been shown to increase response to analytes.30-33 Rather than just

random roughening through chemical processes, several substrates have been fabricated

with pillar architecture, which allows for a periodic surface roughness to be achieved on a

nanoscale.34-36 This strategy forms the basis for the substrates fabricated herein. To

determine if the observed enhancement was indeed due to the substrate architecture, a

substrate was fabricated without reactive ion etching steps to serve as a control with no

pillars. This substrate was compared to a substrate that received all nanofabrication steps

under identical experimental conditions to the control substrate. Both substrates were

exposed to the off-gassed vapor from neat BT for 10 minutes, followed by immediate

spectral acquisition. As seen in Figure 4.1, a characteristic SERS spectrum for benzenethiol

Page 100: Raman Spectroscopy Substrate Optimization and Method

87

Figure 4.1. SERS spectra obtained from substrates (with and without pillar architecture) that were exposed to the off-gassed vapor from neat benzenethiol for 10 minutes. λex = 785 nm, P = 5 mW, t = 60 s

Page 101: Raman Spectroscopy Substrate Optimization and Method

88

was obtained with the substrate where the pillars have been etched. The absence of signal

in the spectrum obtained from the substrate without pillars demonstrates that the pillar

architecture is responsible for the SERS enhancement, rather than just the presence of the

plasmonic gold surface.

To make use of these substrates for gas capture and detection, a porous silicon oxide

(PSO) coating was explored as a method to concentrate the gas-phase analyte molecules at

the plasmonic surface. Prior to the plasma-enhanced chemical vapor deposition (PECVD)

of the PSO, it was necessary to deposit a 2 nm thick layer of alumina (Al2O3) onto the

surface of the Au-DOP to shield the PECVD instrument from the highly conductive

plasmonic metal. The electromagnetic SERS enhancement phenomenon is distance-

dependent, and for most substrates the signal intensity decays rapidly within a few

nanometers from the surface.37, 38 To investigate any detrimental effect caused by the

alumina deposition, a pair of substrates were prepared from the same nanofabricated wafer;

one with the alumina layer and one without. The resulting substrates were immersed in 1

mg mL-1 benzenethiol (9.08 mM, thiophenol, BT), dispersed in isopropanol (IPA) for 10

minutes, followed by three washes with isopropanol (IPA) to remove any non-adsorbed

BT molecules from the surface. The substrates were transferred to a glass cover slip for

spectral collection (λex = 785 nm, P = 5 mW, t = 30 s), and the resulting spectra (Figure

4.2, below) were normalized to the Si band present at 520 cm-1 prior to peak fitting. The

integrated peak area for each vibrational mode showed an average 40% decrease following

alumina deposition, however usable signal from the analyte was still observed.

Mesoporous silica has been shown to facilitate the transport of small molecules

from solution to the plasmonic surface of a gold nanorod,39 and several porous silicon oxide

Page 102: Raman Spectroscopy Substrate Optimization and Method

89

Figure 4.2. SERS activity for pre- and post-alumina layer deposition substrates. Spectral analysis shows that a 40% decrease in signal, but the signal obtained after deposition is still reasonable for analysis. λex = 785 nm, P = 5 mW, t = 30 s

Page 103: Raman Spectroscopy Substrate Optimization and Method

90

substrates have been developed for gas sensing.40, 41 The network of pores in the oxide

serves to both funnel the analyte molecules toward the plasmonic surface and to impede

the exodus of molecules from that surface in our fabricated substrates. First, the thickness

of the PSO layer was optimized by deposition of layers with 150 – 1000 nm thicknesses

and then exposing the substrates to the vapor off-gassed from a pool of neat BT liquid. The

substrates were kept in contact with the vapors for a period of 10 minutes, followed by 60

s spectral acquisition. The results of these experiments from 300 – 1000 nm are shown in

Figure 4.3. After spectral processing, it was found that the 1000 nm (1 μm) PSO thickness

provided the best enhancement across the entire spectral region of interest, and this

thickness was used for all subsequent experiments.

The optimized substrates were subsequently tested with the common Raman

reporter molecules neat BT and neat pyridine using the off-gassing protocol. The exposure

time was varied with each analyte to ascertain the minimum amount of time required for

vapors to be adsorbed onto the substrate. Exposure times were varied from 10 minutes

down to 30 seconds for each analyte (Figure 4.4). For these strongly adsorbing analytes,

the minimal exposure time (30 s) allowed for successful detection of the analytes, as

evidenced by the characteristic SERS spectra obtained with 1 s collection time. Increasing

exposure time,from 30 s to 5 min results in slight increases in intensity, with maximum

intensity at 10 min exposure time.

After analysis with highly concentrated (~10 M) analytes, it was necessary to

evaluate the substrate performance at concentrations more representative of typical VOC

levels. A 1 M solution of BT in ethanol (EtOH) was prepared, then subsequently diluted to

Page 104: Raman Spectroscopy Substrate Optimization and Method

91

Figure 4.3. SERS spectra obtained from 10-minute exposure time to off-gassed neat benzenethiol vapor, demonstrating the best enhancement for the substrate occurs with 1000 nm of porous silicon oxide (PSO). λex = 785 nm, P = 5 mW, t = 60 s

Page 105: Raman Spectroscopy Substrate Optimization and Method

92

Figure 4.4. SERS spectra obtained after off-gassing a) neat benzenethiol and b) neat pyridine onto optimized Au disc-on-pillar arrays with 1000 nm of porous silicon oxide demonstrating that it is possible to rapidly detect either analyte with as little as 30 s of exposure to the vapor. λex = 785 nm, P = 5 mW, t = 1 s

Page 106: Raman Spectroscopy Substrate Optimization and Method

93

100, 10, and 1 ppm in water (909, 90.9, and 9.09 μM, respectively). The resulting ppm

solutions (1 mL total volume) were transferred to glass vials, and the fabricated substrates

were suspended in the headspace above the liquid. The substrates were kept in contact with

the vapor off-gassed from the aqueous BT solutions for 10 minutes prior to 10 s spectral

collection. The spectra were post-processed using GRAMS/AI software

(ThermoScientific), which included spectral subtraction of the unexposed substrate to

remove the contribution of the substrate itself, followed by baseline correction, second-

order polynomial Savitsky-Golay smoothing, offset correction, and peak fitting to generate

integrated intensity values.. As shown in Figure 4.5, it is possible to obtain Raman

scattering signal from part-per-million concentrations of BT with relatively short (10

minute) exposure times to the vapor.

The integrated intensity values were then plotted as a function of concentration on

a logarithmic scale, and a linear regression analysis was performed. For the peak located

at 1572 cm-1, the R2 value obtained from the linear regression was determined to be 0.99593

(Figure 4.6). The linear regressions for the peaks at 996, 1020, and 1070 cm-1 show similar

R2 values (0.83645 to 0.90388; see Figure 4.7). To determine the mean noise obtained from

the substrate, three silent regions of each of the ppm spectra were analyzed: 1200 – 1400,

1500 – 1525, and 1680 – 1710 cm-1. The noise values at each pixel (n = 209) were averaged

to yield a mean noise (N) value of 272 counts. A signal value of 3N was substituted into

the 1572 cm-1 equation, which showed the most linear relationship in the regression

analysis, to determine the LOD of 0.234 ppm, or 234 parts-per-billion (ppb).

Page 107: Raman Spectroscopy Substrate Optimization and Method

94

Figure 4.5. SERS spectra obtained after 10 minutes of off gassing from a) 100 ppm, b) 10 ppm, and c) 1 ppm benzenethiol in water. λex = 785 nm, P = 5 mW, t = 10 s

Page 108: Raman Spectroscopy Substrate Optimization and Method

95

Figure 4.6. Linear regression analysis of integrated intensity of the 1572 cm-1 peak obtained from off gassing aqueous parts-per-million solutions of benzenethiol.

Page 109: Raman Spectroscopy Substrate Optimization and Method

96

Figure 4.7. Linear regression analysis of the (a) 996, (b) 1020, (c) 1070, and (d) 1572 cm-1 peaks obtained from the off-gassing of parts-per-million aqueous solutions of benzenethiol.

Page 110: Raman Spectroscopy Substrate Optimization and Method

97

Due to the presence of lone electron pairs on the nitrogen of pyridine and sulfur of

benzenethiol, these molecules readily adsorb onto the gold surface of our SERS substrates

through donation of a member of the lone pair into the half-filled 5d orbital of the gold

atoms on the surface. These electronic interactions make these two molecules ideal for

characterization of a wide variety of SERS substrates from colloids to other multi-

dimensional substrates. To fully assess the Au-DOP array’s capability for gas detection,

two non-adsorbing analyte molecules were chosen: benzene and ortho-xylene. Because of

the electronic stability conferred by the aromaticity of these two molecules, there is little

electronic motivation for strong adsorption onto the plasmonic metal through donation of

electrons to the Au atom, and the molecules are not strongly bound to the surface. When

using the previously established off-gassing methods, spectra were not obtained within the

short timescales previously employed. To obtain spectra of these molecules, the substrates

were kept in contact with the vapor overnight, and the substrates were quickly transferred

to the confocal Raman system for spectral acquisition (Figure 4.8).

The spectrum of the unexposed Au-DOP array was collected to serve as a baseline

(data not shown), and then the array was exposed to the off-gassed vapors from the neat

analyte. For the array exposed to benzene (Figure 4.8a), variation in the portion of the array

that was analyzed prior to exposure causes spectral artifacts to be present at ~700 cm-1.

However, the characteristically intense benzene ring breathing mode is found at 990 cm-1,

which is slightly downshifted from the normal position of 992 cm-1. When conducting

SERS experiments, it is common to see vibrational modes shifted by 1-30 cm-1 from their

Page 111: Raman Spectroscopy Substrate Optimization and Method

98

Figure 4.8. SERS spectra obtained after Au-DOP substrate was kept in contact with vapors off-gassed from pools of a) neat benzene and b) ortho-xylene. λex = 785 nm, P = 5 mW, t = 60 s

Page 112: Raman Spectroscopy Substrate Optimization and Method

99

positions in the normal Raman spectrum.42, 43 The Raman cross-section of the benzene ring

breathing mode demonstrates an inverse relationship with excitation wavelength,44 and it

is possible that the substrate could be optimized for shorter wavelength excitations to allow

for easier detection of benzene vapor when needed. The peaks observed for ortho-xylene

(Figure 4.8b; 516, 579, 731, 1049, and 1220 cm-1) are reasonably shifted from the normal

Raman vibrational modes (508, 585, 737, 1055, and 1225 cm-1).45

In the non-adsorbing analyte case, the substrates required a longer exposure to the

vapor before signal was observed. Because the analyte molecules have no electronic

motivation to adsorb to the plasmonic metal substrate, a majority of analyte molecules are

believed to be swept out of plasmonic hot spots by simply moving the substrate through

the air. This can be corrected by keeping the substrate in contact with the vapors during

spectral acquisition. Normal Raman spectra of molecules in the gas phase are difficult to

obtain due to low particle densities present in gaseous analytes, and as such any observed

signal obtained by enclosing the SERS substrate within a cell with the gas of interest would

be from molecules that are present at the surface of the plasmonic metal. This implies that

the Au-DOP array can be used for process monitoring, where a gas stream can be diverted

to a cell with a SERS substrate present prior to collection of the Raman spectrum.

Page 113: Raman Spectroscopy Substrate Optimization and Method

100

4.5 Conclusions

In summary, we have developed a SERS substrate that can detect off-gassed vapors

from several volatile organic compounds. Readily adsorbing molecules, including

benzenethiol and pyridine, can be detected with as little as 30 s of vapor contact and

spectral acquisitions of 1 s. Detection of neat benzenethiol vapor has been demonstrated at

a 1 ppm level, with LOD determined to be 234 ppb. For non-adsorbing analytes, exposure

times are lengthened due to weaker interactions between the substrate and the analyte

molecule, but the substrate shows promise as a means for monitoring of gaseous reaction

by-products in an on-line setting.

4.6 Acknowledgements

The authors would like to thank the University of Tennessee, Knoxville for start-up

funding. A portion of this research was conducted at the Center for Nanophase Materials

Sciences (CNMS) at Oak Ridge National Laboratory, which is a DOE Office of Science

User Facility, under CNMS User Proposal #CNMS2017- 073.

Page 114: Raman Spectroscopy Substrate Optimization and Method

101

4.7 References

1. U. S. Environmental Protection Agency. Volatile Organic Compounds' Impact on

Indoor Air Quality. https://www.epa.gov/indoor-air-quality-iaq/volatile-organic-

compounds-impact-indoor-air-quality#Health_Effects (accessed April 11, 2019).

2. Huff, J., Benzene-induced Cancers: Abridged History and Occupational Health

Impact. Int. J. Occup. Environ. Health 2013, 13 (2), 213-221.

3. Nakhleh, M. K.; Amal, H.; Jeries, R.; Broza, Y. Y.; Aboud, M.; Gharra, A.;

Ivgi, H.; Khatib, S.; Badarneh, S.; Har-Shai, L.; Glass-Marmor, L.; Lejbkowicz, I.;

Miller, A.; Badarny, S.; Winer, R.; Finberg, J.; Cohen-Kaminsky, S.; Perros, F.;

Montani, D.; Girerd, B.; Garcia, G.; Simonneau, G.; Nakhoul, F.; Baram, S.; Salim,

R.; Hakim, M.; Gruber, M.; Ronen, O.; Marshak, T.; Doweck, I.; Nativ, O.; Bahouth,

Z.; Shi, D.-y.; Zhang, W.; Hua, Q.-l.; Pan, Y.-y.; Tao, L.; Liu, H.; Karban, A.;

Koifman, E.; Rainis, T.; Skapars, R.; Sivins, A.; Ancans, G.; Liepniece-Karele, I.;

Kikuste, I.; Lasina, I.; Tolmanis, I.; Johnson, D.; Millstone, S. Z.; Fulton, J.; Wells, J.

W.; Wilf, L. H.; Humbert, M.; Leja, M.; Peled, N.; Haick, H., Diagnosis and

Classification of 17 Diseases from 1404 Subjects via Pattern Analysis of Exhaled

Molecules. ACS Nano 2016, 11 (1), 112-125.

4. Adamu, A. I.; Ozturk, F. E.; Bayindir, M., Binary coded identification of

industrial chemical vapors with an optofluidic nose. Appl. Opt. 2016, 55 (36).

5. Kamal, M. S.; Razzak, S. A.; Hossain, M. M., Catalytic oxidation of volatile

organic compounds (VOCs) – A review. Atmos. Environ. 2016, 140, 117-134.

Page 115: Raman Spectroscopy Substrate Optimization and Method

102

6. Guo, H.; Ling, Z. H.; Cheng, H. R.; Simpson, I. J.; Lyu, X. P.; Wang, X. M.;

Shao, M.; Lu, H. X.; Ayoko, G.; Zhang, Y. L.; Saunders, S. M.; Lam, S. H. M.;

Wang, J. L.; Blake, D. R., Tropospheric volatile organic compounds in China. Sci. Total

Environ. 2017, 574, 1021-1043.

7. Zhang, S.; You, J.; Kennes, C.; Cheng, Z.; Ye, J.; Chen, D.; Chen, J.; Wang,

L., Current advances of VOCs degradation by bioelectrochemical systems: A review.

Chem. Eng. J. 2018, 334, 2625-2637.

8. Panigrahi, S.; Sankaran, S.; Mallik, S.; Gaddam, B.; Hanson, A. A., Olfactory

receptor-based polypeptide sensor for acetic acid VOC detection. Mater. Sci. Eng. C

2012, 32 (6), 1307-1313.

9. Reinnig, M.-C.; Müller, L.; Warnke, J.; Hoffmann, T., Characterization of

selected organic compound classes in secondary organic aerosol from biogenic VOCs by

HPLC/MS n. Anal. Bioanal. Chem. 2008, 391 (1), 171-182.

10. Blakeman, K. H.; Cavanaugh, C. A.; Gilliland, W. M.; Ramsey, J. M., High

pressure mass spectrometry of volatile organic compounds with ambient air buffer gas.

Rapid Commun. Mass Spectrom. 2017, 31 (1), 27-32.

11. Laskin, J.; Laskin, A.; Nizkorodov, S. A., Mass Spectrometry Analysis in

Atmospheric Chemistry. Anal. Chem. 2017, 90 (1), 166-189.

12. Xu, K.; Fu, C.; Gao, Z.; Wei, F.; Ying, Y.; Xu, C.; Fu, G., Nanomaterial-based

gas sensors: A review. Instrum. Sci. Technol. 2017, 46 (2), 115-145.

13. Buszewski, B.; Rudnicka, J.; Ligor, T.; Walczak, M.; Jezierski, T.; Amann, A.,

Analytical and unconventional methods of cancer detection using odor. TrAC, Trends

Anal. Chem. 2012, 38, 1-12.

Page 116: Raman Spectroscopy Substrate Optimization and Method

103

14. Qian, C.; Guo, Q. H.; Xu, M. M.; Yuan, Y. X.; Yao, J. L., Improving the SERS

detection sensitivity of aromatic molecules by a PDMS-coated Au nanoparticle

monolayer film. RSC Adv. 2015, 5 (66), 53306-53312.

15. Zhao, C.; Gan, X.; Yuan, Q.; Hu, S.; Fang, L.; Zhao, J., High-Performance

Volatile Organic Compounds Microsensor Based on Few-Layer MoS2-Coated Photonic

Crystal Cavity. Adv. Opt. Mater. 2018.

16. Kneipp, K.; Kneipp, H.; Itzkan, I.; Dasari, R. R.; Feld, M. S., Ultrasensitive

Chemical Analysis by Raman Spectroscopy. Chem. Rev. 1999, (99), 2957-2975.

17. U. S. Occupational Safety & Health Administration. Permissible Exposure Limits

– Annotated Tables. https://www.osha.gov/dsg/annotated-pels/index.html (accessed May

15).

18. Masango, S. S.; Hackler, R. A.; Large, N.; Henry, A.-I.; McAnally, M. O.;

Schatz, G. C.; Stair, P. C.; Van Duyne, R. P., High-Resolution Distance Dependence

Study of Surface-Enhanced Raman Scattering Enabled by Atomic Layer Deposition.

Nano Lett. 2016, 16 (7), 4251-4259.

19. Camden, J. P.; Dieringer, J. A.; Wang, Y.; Masiello, D. J.; Marks, L. D.;

Schatz, G. C.; Van Duyne, R. P., Probing the structure of single-molecule surface-

enhanced Raman scattering hot spots. J. Am. Chem. Soc. 2008, 130 (38), 12616-7.

20. Sharma, B.; Frontiera, R. R.; Henry, A. I.; Ringe, E.; Van Duyne, R. P., SERS:

Materials, applications, and the future. Mater. Today 2012, 15 (1-2), 16-25.

21. Piorek, B. D.; Lee, S. J.; Moskovits, M.; Meinhart, C. D., Free-surface

microfluidics/surface-enhanced Raman spectroscopy for real-time trace vapor detection

of explosives. Anal. Chem. 2012, 84 (22), 9700-5.

Page 117: Raman Spectroscopy Substrate Optimization and Method

104

22. Zapata, F.; Lopez-Lopez, M.; Garcia-Ruiz, C., Detection and identification of

explosives by surface enhanced Raman scattering. Appl. Spectrosc. Rev. 2016, 51 (3),

207-242.

23. Qiao, X.; Su, B.; Liu, C.; Song, Q.; Luo, D.; Mo, G.; Wang, T., Selective

Surface Enhanced Raman Scattering for Quantitative Detection of Lung Cancer

Biomarkers in Superparticle@MOF Structure. Adv. Mater. 2017.

24. Mosier-Boss, P. A., Review of SERS Substrates for Chemical Sensing.

Nanomaterials (Basel) 2017, 7 (6).

25. Mosier-Boss, P. A.; Lieberman, S. H., Detection of volatile organic compounds

using surface enhanced Raman spectroscopy substrates mounted on a thermoelectric

cooler. Anal. Chim. Acta 2003, 488 (1), 15-23.

26. Oh, M. K.; De, R.; Yim, S. Y., Highly sensitive VOC gas sensor employing deep

cooling of SERS film. J. Raman Spectrosc. 2018, 49 (5), 800-809.

27. Chou, A.; Jaatinen, E.; Buividas, R.; Seniutinas, G.; Juodkazis, S.; Izake, E. L.;

Fredericks, P. M., SERS substrate for detection of explosives. Nanoscale 2012, 4 (23),

7419-24.

28. Oh, M.-K.; De, R.; Yim, S.-Y., Highly sensitive VOC gas sensor employing deep

cooling of SERS film. J. Raman Spectrosc. 2018, 49 (5), 800-809.

29. Agapov, R. L.; Srijanto, B.; Fowler, C.; Briggs, D.; Lavrik, N. V.; Sepaniak, M.

J., Lithography-free approach to highly efficient, scalable SERS substrates based on

disordered clusters of disc-on-pillar structures. Nanotechnology 2013, 24 (50).

Page 118: Raman Spectroscopy Substrate Optimization and Method

105

30. Macias, G.; Alba, M.; Marsal, L. F.; Mihi, A., Surface roughness boosts the

SERS performance of imprinted plasmonic architectures. J. Mater. Chem. C 2016, 4 (18),

3970-3975.

31. Sur, U. K., Surface-Enhanced Raman Scattering. In Raman Spectroscopy and

Applications, 2017.

32. Wijesuriya, S.; Burugapalli, K.; Mackay, R.; Ajaezi, G.; Balachandran, W.,

Chemically Roughened Solid Silver: A Simple, Robust and Broadband SERS Substrate.

Sensors 2016, 16 (10).

33. Yang, S.; Slotcavage, D.; Mai, J. D.; Guo, F.; Li, S.; Zhao, Y.; Lei, Y.;

Cameron, C. E.; Huang, T. J., Electrochemically created highly surface roughened Ag

nanoplate arrays for SERS biosensing applications. J. Mater. Chem. C 2014, 2 (39),

8350-8356.

34. Huang, Z.; Meng, G.; Huang, Q.; Yang, Y.; Zhu, C.; Tang, C., Improved SERS

Performance from Au Nanopillar Arrays by Abridging the Pillar Tip Spacing by Ag

Sputtering. Adv. Mater. 2010, 22 (37), 4136-4139.

35. Magno, G.; Bélier, B.; Barbillon, G., Al/Si Nanopillars as Very Sensitive SERS

Substrates. Materials 2018, 11 (9).

36. Schmidt, M. S.; Hübner, J.; Boisen, A., Large Area Fabrication of Leaning

Silicon Nanopillars for Surface Enhanced Raman Spectroscopy. Adv. Mater. 2012, 24

(10), OP11-OP18.

37. Dieringer, J. A.; McFarland, A. D.; Shah, N. C.; Stuart, D. A.; Whitney, A. V.;

Yonzon, C. R.; Young, M. A.; Zhang, X. Y.; Van Duyne, R. P., Surface enhanced

Page 119: Raman Spectroscopy Substrate Optimization and Method

106

Raman spectroscopy: new materials, concepts, characterization tools, and applications.

Faraday Discuss. 2006, 132, 9-26.

38. Masango, S. S.; Hackler, R. A.; Large, N.; Henry, A. I.; McAnally, M. O.;

Schatz, G. C.; Stair, P. C.; Van Duyne, R. P., High-Resolution Distance Dependence

Study of Surface-Enhanced Raman Scattering Enabled by Atomic Layer Deposition.

Nano Lett 2016, 16 (7), 4251-4259.

39. Gao, Z.; Burrows, N. D.; Valley, N. A.; Schatz, G. C.; Murphy, C. J.; Haynes,

C. L., In solution SERS sensing using mesoporous silica-coated gold nanorods. Analyst

2016, 141 (17), 5088-5095.

40. Tiemann, M., Porous Metal Oxides as Gas Sensors. Chem. Eur. J. 2007, 13 (30),

8376-8388.

41. Wagner, T.; Haffer, S.; Weinberger, C.; Klaus, D.; Tiemann, M., Mesoporous

materials as gas sensors. Chem. Soc. Rev. 2013, 42 (9), 4036-4053.

42. Sharma, B.; Cardinal, M. F.; Ross, M. B.; Zrimsek, A. B.; Bykov, S. V.;

Punihaole, D.; Asher, S. A.; Schatz, G. C.; Van Duyne, R. P., Aluminum Film-Over-

Nanosphere Substrates for Deep-UV Surface-Enhanced Resonance Raman Spectroscopy.

Nano Lett 2016, 16 (12), 7968-7973.

43. Stacy, A. A.; Van Duyne, R. P., Surface enhanced raman and resonance raman

spectroscopy in a non-aqueous electrochemical environment: Tris(2,2′-

bipyridine)ruthenium(II) adsorbed on silver from acetonitrile. Chem. Phys. Lett. 1983,

102 (4), 365-370.

44. Schomacker, K. T.; Delaney, J. K.; Champion, P. M., Measurements of the

Absolute Raman Cross-Sections of Benzene. J. Chem. Phys. 1986, 85 (8), 4240-4247.

Page 120: Raman Spectroscopy Substrate Optimization and Method

107

45. Moreau, J.; Rinnert, E., Fast identification and quantification of BTEX coupling

by Raman spectrometry and chemometrics. Analyst 2015, 140 (10), 3535-3542.

Page 121: Raman Spectroscopy Substrate Optimization and Method

108

Chapter 5

Normal Raman Analysis of the Composition of Oxidized

Polyacrylonitrile Carbon Fibers

Page 122: Raman Spectroscopy Substrate Optimization and Method

109

This chapter describes the statistical analysis using Raman spectroscopy of a series of 11

oxidized PAN-based carbon fibers prepared via a proprietary method. Low standard

deviations are obtained from the statistical analysis, indicating that the fiber samples are

reasonably uniform along the length of the fiber. Since Raman spectroscopy is a surface-

selective technique, the fibers were further analyzed by cross-sectional rastering and

mapping, showing that several of the fibers have core-sheath microstructure, but with

varying placement of the core.

This chapter is in preparation for submission in Spring 2020.

Page 123: Raman Spectroscopy Substrate Optimization and Method

110

5.1 Abstract

The surface and internal structure of a series of carbon fibers made from oxidized-

PAN precursors was investigated using Raman spectroscopy. First, the surface properties

were investigated using statistical analysis; for each fiber, five spots were selected at

random along the length of the fiber. Three Raman spectra were collected at each of these

spots (λex = 488 nm, P = 2 mW, t = 60 s), and the resulting spectra were peak fitted using

a novel peak fitting modality that accounts for structural features present in the Raman

spectra. For each spectrum at each of these spots, the integrated intensity ratio of the D-

and G-band was calculated and the crystallite size (La) was calculated from this ratio. For

the ensemble of 11 fiber samples, the ID/IG ratios ranged from 0.974 – 1.166, with standard

deviations from 0.065 – 0.112 (6.1 – 10.8% relative standard deviation), and the values

obtained for the La of each of these fibers ranged from 11.8 – 14.1 nm, with standard

deviations of 0.8 – 1.4 nm (5.9 – 10.2% relative standard deviation). Due to the low RSD,

these fibers are very reproducibly oxidized at the surface. To further investigate the

internal structure of the fibers, cross-sectional rastering and mapping strategies were

employed, and the results confirmed various degrees of core-sheath microstructure on the

interior of the fibers. This core-sheath morphology is observed in electron microscopy and

is confirmed spectroscopically herein. For some of the fibers, the core is well-defined and

centrally located, while for others the core is located primarily to one side, or in multiple

locations. The use of Raman spectroscopic methodology provides insight into the internal

structure of the fibers and can be used in the future to explain physical properties of these

fiber samples.

Page 124: Raman Spectroscopy Substrate Optimization and Method

111

5.2 Introduction

Carbon fibers (CFs) bear a unique set of properties, including high mechanical

properties, light weight, high chemical and thermal resistance, and excellent electrical and

thermal conductivity.1 Common types of precursors are used in the fabrication of CFs

include gaseous precursors, rayon, pitch, and polyacrylonitrile (PAN).1, 2 For the past three

decades, carbon fibers have been primarily produced from a PAN precursor, due in large

part to the low carbon yields and tensile strength of rayon-based CFs and the inferior

mechanical properties of mesophase pitch-based CFs. 3, 4 PAN precursors typically go

through a series of treatment steps, including stabilization, carbonization, and

graphitization; typically the fiber is stretched and oxidized at a temperature between 200-

300 °C to cyclize the fiber into stable intermediate.5 The stabilized fiber then goes through

high temperature treatment to carbonize the stabilized fiber, followed by graphitization.

Raman spectroscopy is a powerful analytical technique that has been used to study

a wide variety of sp2 carbon nanomaterials, including graphene,6, 7 graphite,6, 8, 9 carbon

nanotubes,10, 11 and fullerenes.12, 13 In pristine graphene samples, the G-band, which arises

from doubly degenerate E2g vibrational modes, is typically located at ~1580 cm-1 and is the

only observed feature in the first-order Raman region (<1800 cm-1).14, 15 This vibrational

mode is characteristic for networks of sp2 carbon atoms. For less ordered graphitic

materials, the introduction of disorder into the sample gives rise to another first-order

Raman band known as the D-band, which arises from graphene edges and sub-domain

boundaries, that is located at ~ 1350 cm-1.7 To characterize defect quantity in graphitic

materials, the ratio of the D and G bands is commonly used.14

Page 125: Raman Spectroscopy Substrate Optimization and Method

112

Raman spectra show high sensitivity to microstructural changes in carbon materials,

allowing the technique to be used for characterization of carbon-based materials by

elucidating structural properties like stacking order and crystallite size.14, 16 Intially, it was

believed that an empirical relation between intensity ratios of the D- to G-bands developed

by Tuinstra and Koenig17 held true without restriction on crystallite size, but as time has

progressed it is apparent that this relation holds true only for some samples.18 Subsequently,

it was demonstrated by Mernagh et al. that the Raman shift of the D- and G-bands is

excitation wavelength dependent,19 which demonstrated the need for a general equation

that could be used to calculate crystallite size from Raman band intensity at any laser

excitation wavelength. In 2006, Cançado et al. determined the crystallite size (La) of several

nanographite samples obtained from the heat-treatment of diamond-like carbon films

through X-ray diffraction experiments, and then developed an empirical relationship by

using Raman spectroscopy at excitation wavelengths from 457.9 – 647 nm, which gave

rise to the much-needed wavelength dependent equation for determination of crystallite

size (Equation 5.1). 20

𝐿𝐿𝑎𝑎 (𝑛𝑛𝑚𝑚) =560𝐸𝐸𝑙𝑙4

�𝐼𝐼𝐷𝐷𝐼𝐼𝐺𝐺�−1

5.1

where El is the excitation energy of the laser in units of eV, ID is the integrated intensity of

the D band, and IG is the integrated intensity of the G-band. However, considering the laser

wavelength instead, which is inversely proportional to the energy, the more commonly

used Equation 5.2 is generated.

Page 126: Raman Spectroscopy Substrate Optimization and Method

113

𝐿𝐿𝑎𝑎 (𝑛𝑛𝑚𝑚) = (2.4 × 10−10)𝑚𝑚𝑙𝑙4 �𝐼𝐼𝐷𝐷𝐼𝐼𝐺𝐺�−1

5.2

where λl is the excitation wavelength of the laser in nm units.

Herein, we describe the use of Raman spectroscopy to analyze the bulk behavior of

a series of 11 oxidized-PAN based carbon fiber samples. Statistical analysis was performed

(n = 15) on each fiber to determine the mean ID/IG ratio for each fiber along the length, and

the mean La value was calculated using Equation 5.2. The fibers were then analyzed in the

cross-section, using either a rastering technique across the center of an isolated fiber, or

mapping of the entire fiber cross-section. The resulting ID/IG ratios were then plotted as

both heat maps and contour plots to show the microstructure of the interior of each fiber.

5.3 Methods

5.3.1 Samples

A total of 11 samples were obtained from the Penumadu group in the Department

of Civil and Environmental Engineering (Tickle College of Engineering, University of

Tennessee Knoxville). The samples provided were a commercial standard (1), and then a

series of fibers prepared by proprietary methods. The proprietary fibers are referenced

herein as 2, 3, 4, 5, 6, 7, 8, 9, 10, and 11.

5.3.2 Raman Instrumentation

Raman spectra were collected using a home-built confocal Raman system (Figure

5.1) equipped with a 488 nm optically pumped semiconductor laser (Coherent, Santa Clara,

CA). The laser output was aligned through a series of irises (I1 and I2) and focusing lenses

Page 127: Raman Spectroscopy Substrate Optimization and Method

114

Figure 5.1. Schematic of the Raman apparatus used for the analysis of the oxidized PAN fibers.

Page 128: Raman Spectroscopy Substrate Optimization and Method

115

(FL1 and FL2) before entering a series of periscope optics to direct the light into the back

of an inverted microscope (Nikon Ti-U, Nikon Instruments). Light scattered from the

samples was collected via the microscope objective (40× or 100×, Nikon) and directed

through a dichroic mirror and notch filter (not pictured) prior to entering the slit of an

IsoPlane SCT320 spectrometer equipped with a PIXIS 400 charge-coupled device (CCD)

camera (Princeton Instruments, Princeton, NJ). The spectrometer and CCD were controlled

using LightField software (Princeton Instruments).

5.3.3 Spectral Collection

Tows of each fiber were immobilized with minimal tension using laboratory tape

to affix the sample onto the surface of a microscope slide. The samples were inverted so

that the microscope slide was not between the microscope objective and the fiber. Spectra

were collected for 60 s exposure times using either 2 mW (fiber statistical analysis) or 150

μW (cross sectional fiber rastering and fiber mapping) of power.

5.3.4 Spectral Processing

Post-acquisition data processing was completed using GRAMS/AI software (ver.

9.2, ThermoFisher). First, spectra were baseline corrected to remove fluorescence

background contributions, followed by offset correction. The baseline and offset corrected

spectra were peak fitted to extract the integrated intensity (area) of the D- and G-bands,

which accounts for both the maximum intensity and spectral lineshape of the selected

peaks. During peak fitting, the D-band was fitted using a pure Lorentzian function, while

the G-band was fitted with a hybrid Gaussian-Lorentzian function, which typically

included ~ 10% Lorentzian character. A total of four peaks were used to fit the spectral D-

Page 129: Raman Spectroscopy Substrate Optimization and Method

116

and G-bands, with a small disorder band (D″) placed between the D- and G-bands, and a

Gaussian band placed adjacent to the D-band (pre-D). Previously, oxidized PAN fibers

have been characterized using 3 mixed Gaussian-Lorentzian functions to fit the D-, D″-,

and G- bands;21 however, this peak fitting strategy fails to capture all of the spectral

5.4 Results and Discussion

A small tow of each fiber was secured to a microscope coverslip using adhesive

tape. The tows were secured tautly between the anchor points, but without placing the tow

under strain. Normal Raman spectra (2 mW power, 60 s) of the fiber samples were obtained

and the resulting spectra (Figure 5.2) were processed and peak fitted according to the

methods described above. The obtained spectra show that slight variation in spectral

lineshapes and peak widths are observable from the processed spectra. A sample spectrum

(fiber 1) and the associated peak fitting is shown in Figure 5.3.

For each fiber, 5 spots along the tow were randomly selected for Raman analysis.

At each of these spots, a total of 3 spectral acquisitions were obtained and processed,

resulting in 15 spectra for each fiber type. The resulting integrated intensities of the D- and

G-bands (ID and IG, respectively) were tabulated, and the mean, absolute standard

deviation, and relative standard deviation were calculated on the ID/IG ratio for each fiber

type. For each of the fibers, the mean ID/IG ratios fall between 0.974 and 1.166, with

absolute standard deviations of 0.065 to 0.112, with relative standard deviations on the

order of 6-10%. Lower ID/IG ratios indicate a higher degree of graphitization in the oxidized

Page 130: Raman Spectroscopy Substrate Optimization and Method

117

Figure 5.2. The first-order Raman spectra (1000 – 1800 cm-1) obtained from fibers 1 – 11,

showing differences in the spectral lineshapes and line widths for the fiber samples.

Page 131: Raman Spectroscopy Substrate Optimization and Method

118

Figure 5.3. The Raman spectrum obtained from fiber 1 (a) and the fitting of the D- and G-

band of the fiber (b). The D-band is fitted with a Lorentzian lineshape, whereas the G-band

is fitted using a predominantly Gaussian lineshape.

Page 132: Raman Spectroscopy Substrate Optimization and Method

119

PAN fibers, as the G-band from graphene sp2 hybridized carbons has greater signal than

the disorder-induced D-band. These results are shown in Table 5.1.

A cross section of fiber 11 was prepared by immobilizing the tow of fiber in an

epoxy matrix, followed by microtoming to remove a thin section of the fiber mounted in

epoxy. These sections were mounted onto a glass slide for analysis. Spectra were collected

at 1 μm intervals along the cross-section of multiple RM64 fibers using a 100× objective

(Nikon Instruments, NA 0.5-1.3). For these measurements, the power was reduced to 150

μW to ensure that the spot size didn’t become too large. The spectra are presented in Figure

5.4a and 5.4d, below. The initial spectra were collected off the fiber cross-section to

elucidate changes in the microstructure of the fiber at the edges, as well as in the center of

the samples. The Raman signal increases in intensity until the entirety of the laser spot is

focused onto the fiber. As shown in Figures 5.4b and 5.4e, the initial spots have similar

integrated intensities for the D- and G-bands. As the spectra are collected further into the

fiber, the intensity of the G-band starts to predominate, resulting in lower D/G ratios

(Figures 5.4c and 5.4f). Through the interior of the fiber, this trend continues. Nearing the

far edge of the fiber, the D-band contribution starts to become more predominant, resulting

in increases in the D/G ratio. These results indicate that the defect structure of the fiber

tends to be more superficial, with the interior of the fiber showing greater amounts of

graphitization with less defect.

Upon initial investigation of the cross-sectional rastering, it was decided to map the

entirety of a cross-section of a fiber, rather than just an individual line across the center of

the fiber. Small tows of each fiber were immobilized in an epoxy puck, with the fibers

Page 133: Raman Spectroscopy Substrate Optimization and Method

120

Table 5.1. Statistical analysis of the fiber samples (1 – 11).

Fiber Mean

ID/IG

Std.

Deviation RSD% La (nm) Std. Deviation

RSD%

1 0.977 0.098 10.08 14.07 1.43 10.16

2 1.166 0.106 9.12 11.77 1.10 9.35

3 1.097 0.099 9.06 12.50 1.02 8.18

4 1.003 0.091 9.03 13.67 1.27 9.26

5 1.053 0.107 10.12 13.05 1.31 10.02

6 0.974 0.099 10.22 14.11 1.35 9.54

7 1.056 0.065 6.13 12.94 0.76 5.86

8 1.073 0.104 9.65 12.79 1.18 9.20

9 1.041 0.071 6.82 13.13 0.89 6.78

10 1.053 0.094 8.92 13.01 1.06 8.14

11 1.047 0.112 10.71 13.13 1.34 10.18

Page 134: Raman Spectroscopy Substrate Optimization and Method

121

Figure 5.4. Results of fiber 11 cross-section Raman analysis: (a, d) Raman spectra collected at 1 μm intervals across cross sections of fiber 11. (b, e) The integrated intensity of the D- and G-bands at each collection point. (c, f) The D/G ratio at each collection point.

Page 135: Raman Spectroscopy Substrate Optimization and Method

122

oriented toward the flat surfaces of the puck. The fibers were optically imaged via the 100×

objective to delineate the margins of each fiber and to determine the initial point for

collection. The Raman system used is equipped with a programmable XYZ stage that was

programmed into a 10 x 10 grid for most samples, but fiber 8 required a 12 x 10 grid to

map the entirety of the fiber cross section. The step spacing for each of these configurations

was 2 μm in the x- and y-directions and was selected to minimize the overlap of probed

regions from spot to spot, while still generating a resolved map. The grid points were

moved through manually using the stage control software (Prior Scientific), followed by

spectral collection at each point for 60 s. The obtained spectra were processed using a

custom MATLAB script to baseline the signal to remove fluorescence background

contribution and apply a Savitsky-Golay smoothing filter to the signal prior to peak

integration. Once the pre-processing in MATLAB was complete, the spectra were peak

fitted using GRAMS/AI, to determine the integrated D- and G-band intensities, followed

by calculation of the D/G ratio at each spot in the map. The x, y, and D/G ratio value were

then plotted as both a heat map and a contour plot in Origin software. The contour plots

are smoothed using a thin plate spline (TPS) algorithm, which assumes that all the data

points are distributed on an elastic plate, or spline, that spans the grid points by forming a

2-D surface through constraints at the grid points.

The results of the mapping for fiber 1 are found in Figure 5.5, below, which

demonstrates that this fiber has a core-sheath arrangement in which the middle of the fiber

is disordered, but the exterior of the fiber is more highly graphitized and shows a larger G-

band contribution, leading to an overall lower D/G ratio. The fiber also demonstrates a

Page 136: Raman Spectroscopy Substrate Optimization and Method

123

Figure 5.5. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 1.

Page 137: Raman Spectroscopy Substrate Optimization and Method

124

roughly circular shape, as seen from the mapping of the heat map. For fiber 2 (Figure 5.6),

this core-sheath morphology is not as clearly delineated as in the fiber 1, which is a

commercially produced fiber. Two areas of high disorder exist in the interior of the fiber,

and the more graphitized portion of the fiber is amorphous instead of concentrically located

around the center of the fiber.

For fiber 3, there is a well-defined core, with a high degree of disorder concentrated

located in the center of the fiber. There is also an area in which the increased D/G ratio

“smears” toward one side of the fiber, which can be seen in Figure 5.7. However, for fiber

4, shown in Figure 5.8, the unoxidized core does not have a well-defined shape, and

actually shows two areas with high degree of disorder-induced Raman band signal. This

area of disorder is also located to one side of the characteristic “kidney bean” shape of this

fiber. Fibers 1 and 2 do not show this characteristic shape that is observed in microscopy

of the fibers conducted by the Penumadu group.

Figure 5.9 shows the mapping results obtained from fiber 5, which demonstrates a

well-defined core-sheath morphology with the locally high degree of disorder located in

the approximate center of the fiber. However, this fiber shows that the central core structure

is distorted into an oblong shape and exhibits a varied sheath thickness throughout the

perimeter of the fiber. Additionally, the shape of this fiber is not as expected, and the

diameter of the fiber is also significantly smaller than the previously analyzed fibers.

Interestingly, the fiber shown in Figure 5.10 (fiber 6) shows a return to the pseudo-

circular shape observed in fiber 1, which is the commercially available sample. In contrast

to fiber 1, this fiber does not show the unoxidized core material located in the center of the

Page 138: Raman Spectroscopy Substrate Optimization and Method

125

Figure 5.6. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 2.

Page 139: Raman Spectroscopy Substrate Optimization and Method

126

Figure 5.7. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 3.

Page 140: Raman Spectroscopy Substrate Optimization and Method

127

Figure 5.8. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 4.

Page 141: Raman Spectroscopy Substrate Optimization and Method

128

Figure 5.9. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 5.

Page 142: Raman Spectroscopy Substrate Optimization and Method

129

Figure 5.10. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 6.

Page 143: Raman Spectroscopy Substrate Optimization and Method

130

fiber, but rather it is located in one hemisphere of the cross-section. Further, the values of

the D/G ratio for this fiber are significantly reduced, indicating that the fiber is more

completely oxidized throughout the bulk of the cross section.

Fiber 7 (Figure 5.11) demonstrates interesting morphological characteristics,

showing the expected shape but with a near-lack of core-sheath microstructure in the fiber.

The magnitude of the D/G ratios in this fiber are lower than any of the previously discussed

fibers, showing a higher degree of graphitization. There are three major areas of the fiber,

none of which are centrally located, which show the higher D/G ratio. These three sections

indicate the lack of a single core area of the fiber, showing that the oxygen diffusion

through the fiber is more random than in other fibers.

Figure 5.12 shows the results of the mapping for fiber 8. This fiber is larger than

the previous fibers, and also demonstrates the expected shape. The contour plot of this fiber

is dominated by a large area of unoxidized PAN precursor, which is shifted from the center

of the fiber. There is a lobe that shows a high degree of oxidation on the outside of the

fiber, indicated that the oxygen diffusion into this fiber is not deep during fabrication.

Additionally, the magnitude of the D/G ratios for this fiber are significantly higher than the

other fibers, confirming a lower overall degree of oxidation, and by proxy graphitization,

in this sample.

Fiber 9 also shows a completely disordered microstructure through the mapping

analysis (Figure 5.13). Though the overall range of the D/G ratios is lower than the rest of

the fibers, the existence of a well-defined core-sheath structure is absent for this fiber.

Several regions on the exterior of the fiber cross-section show lower degrees of

Page 144: Raman Spectroscopy Substrate Optimization and Method

131

Figure 5.11. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 7.

Page 145: Raman Spectroscopy Substrate Optimization and Method

132

Figure 5.12. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 8.

Page 146: Raman Spectroscopy Substrate Optimization and Method

133

Figure 5.13. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 9.

Page 147: Raman Spectroscopy Substrate Optimization and Method

134

graphitization, in an arrangement that shows the existence of multiple core-like areas. In

order to have this microstructure, the oxygen responsible for the cyclization reaction must

permeate the bulk of the fiber more easily than in other samples. However, this fiber does

show an overall absence of the expected core-sheath structure via spectroscopic analysis.

Lastly, Figure 5.14 shows the mapping results for fiber 10. This fiber demonstrates

a more reasonable core-sheath structure than fiber 9, but again the core is located off-center

in the cross-section. The magnitude of the D/G ratios is again among the lowest of the fiber

samples. There are also two regions where the oxidation of the fiber is lower than the

surrounding areas, showing a distorted core-sheath structure for this fiber.

5.5 Conclusions

To summarize, herein a peak fitting methodology has been optimized for a series

of proprietary oxidized-PAN based carbon fibers. Due to the Raman spectral features of

these fibers when illuminated with 488 nm excitation light, peak fitting was carried out

using a 4 peak fitting strategy consisting of a fully Gaussian peak, a fully Lorenzian peak,

and two hybrid Gaussian-Lorentzian peaks for the area located in the D- and G-band areas.

Statistical analysis was carried out on 11 fiber samples, with mean D/G ratios and La values

reported for each of the fibers (n = 15 spectra). A cross-sectional rastering was carried out

on fiber, showing the change of the D/G ratio moving through the bulk of the fiber sample.

Using this information, mapping experiments were carried out to spectroscopically

elucidate the microstructure of the oxidized-PAN based carbon fibers.

Page 148: Raman Spectroscopy Substrate Optimization and Method

135

Figure 5.14. Heat map (a) and contour plot (b) obtained from mapping the D/G ratio of fiber 10.

Page 149: Raman Spectroscopy Substrate Optimization and Method

136

5.6 Acknowledgements

The authors would like to thank the Penumadu group at The University of

Tennessee, Knoxville for providing the fiber samples, and would like to thank Grace M.

Sarabia for valuable discussions regarding MATLAB code generation.

Page 150: Raman Spectroscopy Substrate Optimization and Method

137

5.7 References

1. Elagib, T. H. H.; Hassan, E. A. M.; Liu, B. H.; Yu, M. H.; Han, K. Q.,

Carbonization performance of pre-oxidized PAN fibers prepared by microwave heating.

IOP Conference Series: Mater. Sci. Eng. 2019, 634.

2. Shin, H. K.; Park, M.; Kim, H.-Y.; Park, S.-J., An overview of new oxidation

methods for polyacrylonitrile-based carbon fibers. Carbon Lett. 2015, 16 (1), 11-18.

3. Chang, H.; Luo, J.; Liu, H. C.; Zhang, S.; Park, J. G.; Liang, R.; Kumar, S.,

Carbon fibers from polyacrylonitrile/cellulose nanocrystal nanocomposite fibers. Carbon

2019, 145, 764-771.

4. Mishra, Y. K.; Alarifi, I. M.; Khan, W. S.; Asmatulu, R., Synthesis of

electrospun polyacrylonitrile- derived carbon fibers and comparison of properties with

bulk form. PLoS One 2018, 13 (8).

5. Rahaman, M. S. A.; Ismail, A. F.; Mustafa, A., A review of heat treatment on

polyacrylonitrile fiber. Polym. Degrad. Stab. 2007, 92 (8), 1421-1432.

6. Ferrari, A. C., Raman spectroscopy of graphene and graphite: Disorder, electron–

phonon coupling, doping and nonadiabatic effects. Solid State Commun. 2007, 143 (1-2),

47-57.

7. Malard, L. M.; Pimenta, M. A.; Dresselhaus, G.; Dresselhaus, M. S., Raman

spectroscopy in graphene. Phys. Rep. 2009, 473 (5-6), 51-87.

8. Pimenta, M. A.; Dresselhaus, G.; Dresselhaus, M. S.; Cançado, L. G.; Jorio, A.;

Saito, R., Studying disorder in graphite-based systems by Raman spectroscopy. Phys.

Chem. Chem. Phys. 2007, 9 (11), 1276-1290.

Page 151: Raman Spectroscopy Substrate Optimization and Method

138

9. Reich, S.; Ferrari, A.; Robertson, J.; Thomsen, C., Raman spectroscopy of

graphite. Philos. Trans. R. Soc. A 2004, 362 (1824), 2271-2288.

10. Araujo, P. T.; Barbosa Neto, N. M.; Sousa, M. E. S.; Angélica, R. S.; Simões,

S.; Vieira, M. F. G.; Dresselhaus, M. S.; Leite dos Reis, M. A., Multiwall carbon

nanotubes filled with Al4C3: Spectroscopic signatures for electron-phonon coupling due

to doping process. Carbon 2017, 124, 348-356.

11. Saito, R.; Hofmann, M.; Dresselhaus, G.; Jorio, A.; Dresselhaus, M. S., Raman

spectroscopy of graphene and carbon nanotubes. Adv. Phys. 2011, 60 (3), 413-550.

12. Dresselhaus, M. S.; Dresselhaus, G.; Eklund, P. C., Raman Scattering in

Fullerenes. J. Raman Spectrosc. 1996, 27 (3-4), 351-371.

13. Kuzmany, H.; Ferrari, A. C.; Robertson, J.; Pfeiffer, R.; Hulman, M.;

Kramberger, C., Raman spectroscopy of fullerenes and fullerene–nanotube composites.

Philos. Trans. R. Soc. A 2004, 362 (1824), 2375-2406.

14. Pimenta, M. A.; Dresselhaus, G.; Dresselhaus, M. S.; Cancado, L. G.; Jorio, A.;

Saito, R., Studying disorder in graphite-based systems by Raman spectroscopy. Phys.

Chem. Chem. Phys. 2007, 9 (11), 1276-91.

15. Vázquez-Santos, M. B.; Geissler, E.; László, K.; Rouzaud, J.-N.; Martínez-

Alonso, A.; Tascón, J. M. D., Comparative XRD, Raman, and TEM Study on

Graphitization of PBO-Derived Carbon Fibers. J. Phys. Chem. C 2011, 116 (1), 257-268.

16. Wang, Y.; Alsmeyer, D. C.; McCreery, R. L., Raman spectroscopy of carbon

materials: structural basis of observed spectra. Chem. Mater. 1990, 2 (5), 557-563.

17. Tuinstra, F.; Koenig, J. L., Raman Spectrum of Graphite. J. Chem. Phys. 1970, 53

(3), 1126-1130.

Page 152: Raman Spectroscopy Substrate Optimization and Method

139

18. Puech; Kandara; Paredes; Moulin; Weiss, H.; Kundu; Ratel, R.; Plewa;

Pellenq; Monthioux, Analyzing the Raman Spectra of Graphenic Carbon Materials from

Kerogens to Nanotubes: What Type of Information Can Be Extracted from Defect

Bands? C 2019, 5 (4).

19. Mernagh, T. P.; Cooney, R. P.; Johnson, R. A., Raman spectra of Graphon

carbon black. Carbon 1984, 22 (1), 39-42.

20. Cançado, L. G.; Takai, K.; Enoki, T.; Endo, M.; Kim, Y. A.; Mizusaki, H.;

Jorio, A.; Coelho, L. N.; Magalhães-Paniago, R.; Pimenta, M. A., General equation for

the determination of the crystallite size La of nanographite by Raman spectroscopy. Appl.

Phys. Lett. 2006, 88 (16).

21. Kim, C.; Park, S.-H.; Cho, J.-I.; Lee, D.-Y.; Park, T.-J.; Lee, W.-J.; Yang, K.-

S., Raman spectroscopic evaluation of polyacrylonitrile-based carbon nanofibers

prepared by electrospinning. J. Raman Spectrosc. 2004, 35 (11), 928-933.

Page 153: Raman Spectroscopy Substrate Optimization and Method

140

Chapter 6

Conclusions and Future Directions

Page 154: Raman Spectroscopy Substrate Optimization and Method

141

Raman spectroscopic techniques are highly desirable analytical techniques that can

provide a wealth of chemical information from complex samples, including cyclic organic

molecules. SERS was used to detect the presence of cortisol in an ethanolic matrix at

physiologically relevant concentrations using a colloidal silver nanoparticle substrate,

which showed an LOD of 177 nM in ethanol. The technique was also coupled with PCA

and applied to physiologically relevant cortisol concentrations in a biologically mimetic

fluid matrix, resulting in a linear response with respect to cortisol concentration in a more

complex matrix.

SERS was also applied to the detection of gas-phase volatile organic compounds

molecules using a fabricated substrate consisting of plasmonic gold discs situated on top

of non-lithographically etched silicon pillars. A porous silicon oxide coating was used to

provide a method to enhance the interaction of analyte molecules at the plasmonic surface,

and the substrate was optimized for the detection of volatile organic compounds. The

substrate was able to detect four different VOC molecules, and shows detection of

benzenethiol in the vapor phase at parts-per-million concentrations. For non-adsorbing

analytes, longer contact times were needed to detect the analyte molecules using a confocal

micro-Raman system; however, the substrate does show potential to be enclosed in a flow

cell that is coupled to a Raman system for the on-line detection of gaseous VOCs in an

industrial or process setting.

Lastly, the chemical specificity of normal Raman spectroscopy was used to explore

the physical properties of carbon fibers prepared via a novel, proprietary method from

polyacrylonitrile precursor. A peak fitting method was developed that resulted in the ability

to statistically analyze the fibers for bulk surface properties. This method was applied to

Page 155: Raman Spectroscopy Substrate Optimization and Method

142

cross-sectional rastering and mapping experiments to elucidate the microstructure of 11

different fiber samples, showing varying degree of core-sheath structure and morphology

using a high-magnification objective.

After demonstrating the ability of Raman-based techniques to be used for the

detection of cyclic organic molecules, future work in these projects will include the use of

SERS to further investigate the detection of cholesterol-derived molecules in human serum

mimics, allowing the ability to reduce analysis time from traditional analytical methods.

The methods described herein can be applied to other corticosteroid molecules, as well as

other intermediates in the steroidogenesis pathways within human physiology. These

methods can also be employed in the detection of other gaseous VOC molecules in process-

and environmental monitoring strategies. Incorporation of the fabricated substrates into a

flow cell for the continuous monitoring of on-line chemical processes should be explored

as a next step in this project. Lastly, comparison of other physical properties with the

Raman analysis of oxidized PAN-based carbon fibers should be explored to determine the

effects of the microstructure on moduli and shear forces. Analysis of the fibers when placed

under varying degrees of tension will also show how the microstructure of the fibers

responds to increased stress and can be used to optimize the fabrication of the fibers and

their physical properties.

Page 156: Raman Spectroscopy Substrate Optimization and Method

143

VITA

Terence Joshua Riffey-Moore was born in Bristol, Tennessee, in November 1980,

the only son of Terry L. Moore and Diana Barker. He attended Bristol Tennessee High

School, where he was active as a member of the concert and marching bands, and also

served as Editor-in-Chief of the school newspaper. After graduation from THS, Josh

continued on to East Tennessee State University, where he earned a Bachelor of Business

Administration in Management in 2008. During his business studies, Josh had been focused

on earning entry to medical school, but when preparing for the entrance exams, the study

of chemistry stood out to him. He re-enrolled at ETSU to pursue a Bachelor of Science in

Chemistry, which was awarded in 2012. Josh then enrolled in the chemistry graduate

program at ETSU, where he completed his research in the lab of Dr. Marina Roginskaya,

focusing on DNA damage induced by carbonate radicals. After this, Josh worked for a few

semesters as an adjunct faculty member at ETSU, and then enrolled at The University of

Tennessee, Knoxville in Fall 2015 for his Ph. D. studies in Physical Chemistry.

During his graduate studies at UTK, Josh worked under Dr. Bhavya Sharma to use

Raman spectroscopy in a wide variety of ways. Work was completed on detection of

cortisol in a biofluid mimic using surface-enhanced Raman spectroscopy (SERS), as well

as on detection of volatile organic compounds in the gas phase using SERS substrates.

Additional work was carried out using normal Raman spectroscopy to characterize

materials.

Josh served as the Secretary of the Executive Committee for the Association of

Chemistry Graduate Students (ACGS) during 2018, and has assisted with two summer

programs from 2017 – 2019: The Forensics Camp for middle schoolers hosted by the

Page 157: Raman Spectroscopy Substrate Optimization and Method

144

Chemistry department, and the Governor’s School for Science and Engineering. In

addition, Josh has volunteered his time as a reviewer for the 2018 Student Spaceflight

Experiments Program and as a conference volunteer at SciX 2018.

Josh is married to his wonderful husband Beau, and they have an amazing son

Isaiah. All the humans in the house are ruled by two Chihuahuas, Hustler and Gigolo, with

a wealth of personalities.