98
Theoretical Simulations of Optical Rotation and Raman Optical Activity of Molecules in Solution by Parag Mukhopadhyay Department of Chemistry Duke University Date:_______________________ Approved: ___________________________ Prof. David N. Beratan, Supervisor ___________________________ Prof. Weitao Yang ___________________________ Prof. Jie Liu ___________________________ Prof. Terrence G. Oas Dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in the Department of Chemistry in the Graduate School of Duke University 2008

ABSTRACT - Duke University

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: ABSTRACT - Duke University

v

Theoretical Simulations of Optical Rotation and Raman Optical Activity of Molecules in Solution

by

Parag Mukhopadhyay

Department of ChemistryDuke University

Date:_______________________

Approved:

___________________________Prof. David N. Beratan, Supervisor

___________________________Prof. Weitao Yang

___________________________Prof. Jie Liu

___________________________Prof. Terrence G. Oas

Dissertation submitted in partial fulfillment ofthe requirements for the degree of

Doctor of Philosophy in the Department ofChemistry in the Graduate School

of Duke University

2008

Page 2: ABSTRACT - Duke University

ABSTRACT

Theoretical Simulations of Optical rotation and Raman Optical Activity of Molecules in Solution

by

Parag Mukhopadhyay

Department of ChemistryDuke University

Date:_______________________

Approved:

___________________________Prof. David N. Beratan, Supervisor

___________________________Prof. Weitao Yang

___________________________Prof. Jie Liu

___________________________Prof. Terrence G. Oas

An abstract of a dissertation submitted in partialfulfillment of the requirements for the degree of

Doctor of Philosophy in the Department ofChemistry in the Graduate School

of Duke University

2008

Page 3: ABSTRACT - Duke University

Copyright byParag Mukhopadhyay

2008

Page 4: ABSTRACT - Duke University

iv

Abstract

Chiroptical spectroscopic techniques such as optical rotation (OR), electronic circular

dichroism (ECD), and Raman optical activity (ROA) provide the means to probe molecular

dissymmetry, which is the arrangement of atoms that is not superimposable on its mirror image.

However, the application of chiroptical methods for conformational and configurational analysis

of molecules (or molecular stereochemistry) requires the accurate and efficient computation of

chiroptical response properties. Access to a battery of accurate methods, especially for OR, have

just emerged over the last decade. A combination of quantum chemical electronic structure

methods and chiroptical spectroscopy can now be used to establish a comprehensive

understanding of molecular stereochemistry. A key challenge in stereochemical analysis is to

compute the chiroptical response of molecules in solution, which is the focus of our research in

optical activity.

Although the common perception is that chiroptical responses are solely determined by a

chiral solute’s electronic structure in its environment, we demonstrate that chiral imprinting

effects on media, and molecular assembly effects can dominate chiroptical signatures. Both these

effects are strongly influenced by intermolecular interactions that are essential to accurately

describe chiroptical signatures of molecules. Chiroptical response of molecules spans a range of

large positive and negative values, and hence, the measured chiroptical response represents an

ensemble average of orientations. Thus, accurate prediction of chiroptical properties requires

adequate conformational averaging. Here we show that OR and ROA of molecules in solution

can be modeled using a combination of structure sampling and electronic structure methods.

Method for computing chiroptical response in the condensed phase using continuum solvation

models and point charge solvent model, although more efficient than explicit solvent treatments,

are often inadequate for describing induced chirality effects. The lack of a priori knowledge of

Page 5: ABSTRACT - Duke University

v

solvent-solute interactions and their influence on the chiroptical signature demands exploration of

thermally averaged solute-solvent clusters and comparison with simpler solvent studies.

Page 6: ABSTRACT - Duke University

vi

Contents

Abstract ................................................................................................................................................. iv

Contents ................................................................................................................................................ vi

List of Tables ...................................................................................................................................... viii

List of Figures....................................................................................................................................... ix

List of Abbreviations ........................................................................................................................... xii

Acknowledgements............................................................................................................................. xiii

1 Optical signature of molecular dissymmetry: Combining theory with experiments to addressstereochemical puzzles ........................................................................................................................ 14

1.1 Overview............................................................................................................................ 14

1.2 Introduction........................................................................................................................ 14

1.3 Molecular dissymmetry and the electric dipole-magnetic dipole polarizability tensor .. 15

1.4 Utilizing the G′(ω) tensor to probe molecular dissymmetry............................................ 19

1.4.1 Configuration and conformation analysis of molecules using chiropticalspectroscopy................................................................................................................................ 19

1.4.2 Chiroptical signatures of intra- and inter-molecular interactions. .......................... 21

1.5 Summary............................................................................................................................ 23

2 Solvent effect on optical rotation: A case study of Methyloxirane in water ............................ 25

2.1 Overview............................................................................................................................ 25

2.2 Introduction........................................................................................................................ 25

2.3 Results and Discussion...................................................................................................... 28

2.4 Conclusions........................................................................................................................ 36

2.5 Materials and methods....................................................................................................... 37

2.5.1 Experimental Procedure ........................................................................................... 37

2.5.2 Molecular dynamics simulation setup and procedure............................................. 37

2.5.3 Computation of optical rotation ............................................................................... 38

Page 7: ABSTRACT - Duke University

vii

3 Contribution of a solute’s chiral solvent imprint to optical rotation......................................... 39

3.1 Overview............................................................................................................................ 39

3.2 Introduction........................................................................................................................ 39

3.3 Results and Discussion...................................................................................................... 41

3.4 Conclusions........................................................................................................................ 45

3.5 Materials and Methods ...................................................................................................... 46

3.5.1 Monte Carlo simulation setup and procedure.......................................................... 46

3.5.2 Calculation of optical rotation.................................................................................. 47

4 Characterizing aqueous solution conformations of a peptide backbone using Raman opticalactivity computation ............................................................................................................................ 48

4.1 Overview............................................................................................................................ 48

4.2 Introduction........................................................................................................................ 49

4.3 Results................................................................................................................................ 52

4.3.1 Low-energy conformations of Ala dipeptide in water ............................................ 52

4.3.2 Structures of H2O-Ala dipeptide clusters ................................................................ 54

4.3.3 ROA and Raman spectra of H2O- and D2O- Ala dipeptide clusters....................... 57

4.4 Discussion.......................................................................................................................... 74

4.5 Conclusions........................................................................................................................ 80

4.6 Methods ............................................................................................................................. 81

4.6.1 MC simulations ........................................................................................................ 81

4.6.2 TD-DFT calculations of ROA.................................................................................. 82

5 Epilogue ...................................................................................................................................... 85

Bibliography ........................................................................................................................................ 88

Biography............................................................................................................................................. 98

Page 8: ABSTRACT - Duke University

viii

List of Tables

Table 4-1 The peptide backbone dihedral angles (φ, ψ) of alanine dipeptide-water clusterstructures from the Monte Carlo simulations used in the quantum mechanical energyminimization. ....................................................................................................................................... 56

Page 9: ABSTRACT - Duke University

ix

List of Figures

Figure 2.1 Optical rotatory dispersion of methyloxirane in the gas phase and in water computedusing an implicit solvent model (A) and a combination of implicit and explicit solvent models(B). Calculations used the TD-DFT/BP86 functional and aug-cc-pVDZ basis set. The conductor-like screening model (COSMO) was used as the implicit solvent model. The vertical bars showthe error estimate of the average optical rotation calculated using the blocking method 71.............. 28

Figure 2.2 Radial distribution function of hydrogen atoms for water around the methyloxiraneoxygen atom......................................................................................................................................... 30

Figure 2.3 Optical rotatory dispersion of (S)-methyloxirane in water computed with an increasingnumber of solvent molecules. Calculations used the TD-DFT/BP86 functional and aug-cc-pVDZbasis sets. The conductor-like screening model (COSMO) was used as the implicit solvent model.The error bars were calculated using the blocking method 71. ........................................................... 31

Figure 2.4 Variation in the dipole-length and dipole-velocity form of calculated optical rotationsusing TD-DFT/BP86 correlation-exchange functional. The specific rotation was averaged for1700 structure snapshots from a molecular dynamic simulation to generate the ORD of (S)-methyloxirane in water. The conductor-like screening model (COSMO) was used as the implicitsolvent model. The vertical bars show the error estimate of the average optical rotation calculatedusing the blocking method 71. .............................................................................................................. 32

Figure 2.5 Computed optical rotatory dispersion of (S)-methyloxirane with the first explicithydration shell. Calculations used different combinations of TD-DFT/BP86/B3LYP functionalsand aug-cc-pVDZ/aug-cc-pVTZ basis sets as described in the text. The conductor-like screeningmodel (COSMO) was used as the implicit solvent model. The bars indicate error estimatescomputed using the blocking method 71.............................................................................................. 33

Figure 2.6 Average optical rotation of (S)-methyloxirane at 589 nm, [α]D, as a function of thenumber of MD simulation snapshots. Calculations used the TD-DFT/BP86 functional and aug-cc-pVDZ basis set. A combination of explicit water molecules (within 0.40 nm of the methyloxiraneoxygen atoms) and an implicit solvent model (COSMO) were used in the calculations. The errorestimate of the average optical rotation was calculated using the blocking method 71. .................... 34

Figure 2.7 Distribution of the optical rotation of (S)-methyloxirane at 589 nm, [α]D, as a functionof water molecule positions defined by the angle τ, in the first hydration shell. The insets showstructures with water molecules on the same or opposite sides from the methyl group ofmethyloxirane, defined as the syn- or anti-configurations, respectively. Water on the opposite sidedominated the explicit solvent contribution to [α]D. .......................................................................... 35

Figure 3.1 Optical rotatory dispersion (ORD) of (S) ()- and (R) ()-methyloxirane-benzeneclusters (shown in the MC structure snapshot), and the solvent imprint () computed usingexplicit solvent models. Calculations used TD-DFT/BP86 functional and SVP basis set; and animplicit solvent () based on the COSMO model. The error bars are calculated using theblocking method 71. .............................................................................................................................. 41

Page 10: ABSTRACT - Duke University

x

Figure 3.2 Optical rotatory dispersion (ORD) of (S)-methyloxirane in benzene () and thecomputed ORD of the chiral solvent imprint (;;;). Calculations used varied TD-DFT/BP86/BLYP correlation-exchange functionals with RI-J and SV/SVP/aug-cc-pVDZ basissets. The specific rotation was averaged over 1000 structure snapshots from a Monte-Carlosimulation to generate the ORD of the chiral solvent structure. ........................................................ 43

Figure 3.3 Optical rotatory dispersion of ethylene oxide in benzene computed using explicitsolvent model. Calculations used TD-DFT/BP86 correlation-exchange functionals with RI-J andaug-cc-pVDZ basis set. The specific rotation was averaged over 1000 structure snapshots from aMonte-Carlo simulation to generate the ORD of ethylene oxide and (S)-mehtyloxirane () inbenzene................................................................................................................................................. 44

Figure 3.4 Optical rotatory dispersion of (S)-methyloxirane in alkylbenzenes................................. 45

Figure 4.1 . A. Ball and stick model for the N-acetylalanine-N′-methylamide (Ala dipeptide). Thepeptide backbone structure in Ala dipeptide is defined by φ and ψ dihedral angles. B. Free energydifferences of Ala dipeptide conformations in water computed using Monte Carlo simulations.αR, αL, β, and PPII conformational regions are shown on the Ramachandran map. Figure 4.1Cshows the computed φ and ψ distribution (yellow) resembles the empirically derived φ, ψdistribution of amino acid residues (black) not in either canonical helix or sheet secondarystructures in protein X-ray crystal structures 127. ................................................................................ 53

Figure 4.2 Ramachandran map of φ, ψ pairs for Ala dipeptide-water cluster structures taken fromthe Monte Carlo simulations. Figure 4.2A shows a collection of Ala dipeptide-water clusterstructures (black) that have the same peptide backbone conformation with different waterarrangements. Figure 4.2B shows the distribution of φ, ψ values for the alanine dipeptide-waterclusters optimized using quantum mechanical energy minimization. Green arrows show that eachpeptide conformation, before and after energy minimization, occupies the same region on theRamachandran map. ............................................................................................................................ 55

Figure 4.3 SCP backscattering ROA (A, C) and Raman (B, D) spectra of H2O-Ala dipeptideclusters. The computed spectra shown in blue and red are averaged over a collection of dipeptideconformations from PPII (inset in A) and αR (inset in C) regions of the Ramchandran map. Theexperimental spectra (green; in arbitrary units) are from reference 126.............................................. 58

Figure 4.4 ROA spectra computed using Ala dipeptide-H2O cluster from the low-energy region ofthe Ramachandran map. The computed spectra shown are averaged over a collection of dipeptideconformations from regions D, E, F and G of the Ramchandran map (inset in A). Theexperimental spectrum (in arbitrary units) is from reference 126. ....................................................... 60

Figure 4.5 ROA spectra computed using Ala dipeptide-H2O cluster from the low-energy region ofthe Ramachandran map. The computed spectra shown are averaged over a collection of dipeptideconformations from regions D, E, F and G of the Ramchandran map (inset in A). Theexperimental spectrum (in arbitrary units) is from reference 126. ....................................................... 62

Figure 4.6 SCP backscattering ROA (A, C) and Raman (B, D) spectra of D2O-Ala dipeptideclusters. The computed spectra shown in blue and red are averaged over a collection of dipeptideconformations from PPII (inset in A) and αR (inset in C) regions of the Ramchandran map. Theexperimental spectra (green; in arbitrary units) are from reference 126.............................................. 63

Page 11: ABSTRACT - Duke University

xi

Figure 4.7 ROA spectra computed using Ala dipeptide-D2O cluster from the low-energy region ofthe Ramachandran map. The computed spectra shown are averaged over a collection of dipeptideconformations from regions A, B and C of the Ramchandran map (inset in A). The experimentalspectrum (in arbitrary units) is from reference 126. ............................................................................. 65

Figure 4.8 ROA spectra computed using Ala dipeptide-D2O cluster from the low-energy region ofthe Ramachandran map. The computed spectra shown are averaged over a collection of dipeptideconformations from regions D, E, F and G of the Ramchandran map (inset in A). Theexperimental spectrum (in arbitrary units) is from reference 126. ....................................................... 66

Figure 4.9 SCP backscattering ROA (A, C) and Raman (B, D) spectra of Ala dipeptide in H2O.The computed spectra shown in orange and pink are averaged over a collection of dipeptideconformations from αL (inset in A) and β (inset in C) regions of the Ramchandran map. Theexperimental spectrum (in arbitrary units) is from reference 126. ....................................................... 68

Figure 4.10 SCP backscattering ROA (A, C) and Raman (B, D) spectra of Ala dipeptide in D2O.The computed spectra shown in orange and pink are averaged over a collection of dipeptideconformations from αL (inset in A) and β (inset in C) regions of the Ramchandran map. Theexperimental spectrum (in arbitrary units) is from reference 126. ....................................................... 69

Figure 4.11 ROA (A, C) and Raman (B, D) scattering cross sections contributions of Aladipeptide (red), water (blue), and peptide-water interactions (black) to the total ROA and Ramanscattering cross sections of Ala dipeptide-water clusters (green). These computed spectra areaveraged over a collection of dipeptide conformations from PPII (A, B) and αR (C, D) regions ofthe Ramachandran map. ...................................................................................................................... 70

Figure 4.12 SCP backscattering ROA spectra computed for a representative PPII (φ = -68° andψ = 135°; blue) and a αR conformation of Ala dipeptide (φ = -73° and ψ = −30°; red).Experimental spectra (green; in arbitrary units) in H2O (A, B) and D2O (C, D) are from reference126. ......................................................................................................................................................... 71

Figure 4.13 ROA intensity differences associated with the vibrations in the low wavenumberrange decomposed into contributions from groups of atoms in Ala dipeptide for a PPII (φ = -68°and ψ = 135°; A) and a αR (φ = -73° and ψ = −30°; B) conformation of Ala dipeptide. The groupsof atoms in Ala dipeptide are shown in panel I. Positive and negative ROA intensity differencesare shown as red and yellow circles, respectively. ............................................................................. 73

Page 12: ABSTRACT - Duke University

xii

List of Abbreviations

OR: Optical rotation

ORD: Optical rotatory dispersion

ROA: Raman optical activity

DFT: Density functional theory

TD-DFT: Time dependent density functional theory

MD: Molecular dynamics simulations

MC: Monte Carlo simulations

Page 13: ABSTRACT - Duke University

xiii

Acknowledgements

First and foremost, I thank my advisor Prof. David N. Beratan for his guidance,

encouragement and support throughout my graduate studies. I also thank my committee members

Prof. Weitao Yang, Prof. Jie Liu, and Prof. Terrence G. Oas for insightful discussions. Prof. Oas

also introduced me to the interdisciplinary training program in Structure Biology and Biophysics,

and I express my sincere gratitude for his constant support for my academic and professional

development.

I thank various past and present members of the Beratan group: Dr. Gerard Zuber, Dr.

Michael Rock Goldsmith, Dr. Michael Peterson, Dr. Alexander Balaeff, and Dr. Ilya Balabin, for

insightful discussions.

Last, but not least, I thank my wife Dr. Pooja Arora, my parents and sister for their

support and constant encouragement.

Page 14: ABSTRACT - Duke University

14

1 Optical signature of molecular dissymmetry: Combining

theory with experiments to address stereochemical puzzles

1.1 Overview

Modern chemistry emerged from a quest to understand the rules governing the shapes of

objects at the molecular level. An immediate consequence of van’t Hoff’s description of carbon’s

tetravalency is molecular dissymmetry. Chiroptical spectroscopy, including optical rotatory

dispersion (ORD), electronic circular dichroism (ECD), and Raman optical activity (ROA)

provide means to probe the dissymmetric arrangement of atoms in space. However, deciphering

the molecular shape information encoded in chiroptical response properties requires the accurate

and efficient theoretical simulation of these response properties. Indeed, access to a battery of

accurate theoretical methods has just emerged over the last decade and these methods are forging

strong links between chiroptical spectroscopy and molecular structure. Thus, this chapter briefly

describes the applications of chiroptical spectroscopy to molecular structure analysis, and is

organized as follows: in the introduction we give a brief historical perspective of the advent of

molecular stereochemistry, in 1.3 we review the physical origins of chiroptical phenomena, the

optical activity tensor, G′′(ω), and then in 1.4 we explore the value of linking theory to

experiment in order to correlate chiroptical response to molecular stereochemistry. Finally, in the

summary we briefly describe our research finding in optical activity of molecules in solution.

1.2 Introduction

In 1815 Biot found that plane-polarized light is rotated by a characteristic angle when it

passes through some organic liquids. This behavior became known as optical activity. In 1848

Pasteur separated crystals of the salts of tartaric acid into right- and left-handed hemihedral

crystals. Solutions of these crystals rotated polarized light in opposite directions, leading him to

conclude that molecular dissymmetry results in optical activity 1,2. Molecular dissymmetry is

Page 15: ABSTRACT - Duke University

15

arrangement of atoms in a manner that is not superimposable on their mirror image, a necessary

condition for optical activity. Subsequently, in 1874 van’t Hoff showed that the tetrahedral

geometry of a carbon atom attached to four different atoms (or groups of atoms) is dissymmetric

and hence optically active. During the same time Joseph Achille Le Bel 3, independently showed

that for a molecule of type XY4, where X is an atom or group of atoms joined to four univalent

atoms Y, if the three Ys are replaced by different substituents, then from a purely geometrical

consideration the resulting molecule is dissymmetric and hence optically active. Thus, van’t Hoff

and Le Bel’s investigations of molecular optical activity laid the foundation of modern molecular

stereochemistry 4.

The quantum mechanical (QM) description of optical activity 5-8 followed just a few years

after the emergence of the Schrödinger equation yet, reliable computational methods to predict

OR are just ten years old! 9,10. Following the first accurate ab initio calculation of OR 11, advances

in computational methods for the reliable prediction of OR were reported 9,12. However, the use of

OR for structural analysis of chiral molecules has been limited because of the lack of a direct

intuitive connection between observed rotations and molecular structure. Indeed, no general rules

even exist to predict the sign of the OR for a given structure. Our aim is to compute molecular

chiroptical signatures so that the data may be used in combination with experiments to assign and

interpret the stereochemistry (both conformation and configuration) of the molecule.

1.3 Molecular dissymmetry and the electric dipole-magneticdipole polarizability tensor

Rosenfeld and Condon formulated a modern description for the propagation of polarized

radiation in optically active media 6. Based on a semiclassical theory, they showed that the optical

rotatory parameter β, describes the response of a chiral molecule to polarized radiation. The

specific rotation [α]λ (in deg [dm(g/cm3)]-1) of a chiral molecule is related to its optical rotatory

parameter β:

α[ ]λ =−1.3423×10−4 ×β × ˜ ν 2

3M (1)

Page 16: ABSTRACT - Duke University

16

where M is the molar mass (g/mol),

˜ ν is the radiation frequency in cm-1 and β is in atomic units

(bohr).

The optical rotatory parameter β is an isotropic average derived from the trace of the electric

dipole-magnetic dipole tensor (or the optical activity tensor), G′′(ω):

β = −′ G xx (ω) + ′ G yy (ω) + ′ G zz(ω)

3ω= −

13ω

Tr ′ G (ω)[ ] (2)

G′′(ω) is a second rank tensor and the QM expression for G′′(ω) is:

′ G αβ ω( ) = −4πh

ωω j 0

2 −ω 2j≠0∑ Im 0 ˆ µ α j j ˆ m β 0{ } (3)

where h is Planck’s constant,

0 and

j are the ground and excited states, respectively;

ˆ µ α and

ˆ m β are electric and magnetic dipole moment operators;

ω0 j is the angular transition frequency

(

ω = 2πc ˜ ν ;

˜ ν is frequency in cm-1) and

ω is the angular frequency of the incident radiation. The

rotational strength is

R0 j = −i 0µ j • j m 0{ } (

µ and

m are electric and magnetic transition

moment vectors, respectively) and defines the intensity of ECD bands.

R0 j and β (equation 2) are

related by:

β = −2π3h

1ω0 j2 −ω 2

j≠0∑ R0 j (4)

Thus, OR can be computed from ECD spectrum. In achiral systems Tr [G′′(ω)] cancels. The

optical rotatory parameter β is small because Tr [G′′(ω)] nearly cancels in chiral systems 13. As a

consequence, small changes in the electronic distribution can produce large OR changes. In a QM

calculation, changing basis set, level of electron correlation, solvent, molecular geometry, or

treatment of vibrational contributions can have a large influence on β. Indeed, it remains

challenging to predict OR accurately for chiral molecules. Recently, coupled-cluster (CC) and

time dependent density functional theory (TD-DFT) methods have been used to predict OR 9,

although the computational efficiency of TD-DFT methods favors their use in many OR

calculations 12.

Page 17: ABSTRACT - Duke University

17

ROA measures the intensity difference of Raman scattered right- and left-circularly polarized

light 14. Scattered light originates from the radiation field generated by the oscillating electric and

magnetic multipole moments induced in a molecule by the incident light wave 15. The interference

between light waves scattered via the molecular polarizability (α) and optical activity response

(G′′ and A) yields a dependence of the scattered intensity on the degree of circular polarization of

the incident light and to the circular component on the scattered light, which is the physical basis

of ROA 16. The electric dipole-electric dipole (α) and electric dipole-electric quadrupole (A)

tensors are 15:

ααβ ω( ) =4πh

ω j 0

ω j 02 −ω 2

j≠0∑ Re 0 ˆ µ α j j ˆ µ β 0{ } (5)

Aαβγ ω( ) =4πh

ω j 0

ω j 02 −ω 2

j≠0∑ Re 0 ˆ µ α j j ˆ Θ βγ 0

(6)

where

ˆ Θ βγ is the quadrupole moment operator.

Zuber et al. recently demonstrated accurate and efficient prediction of ROA spectra using

TD-DFT methods 17. The approach avoids calculating the four-center Coulomb integrals and

omits contributions of the electric dipole-electric quadrupole tensor (A) to the ROA differential

scattering cross-section. We used the same scheme in the off-resonance approximation 16 to

calculate the circular difference differential scattering cross-sections per unit solid angle

−Δndσ π( )SCPdΩ

for back scattered circular polarization (SCP) 18 with naturally polarized (n)

incident light:

−Δndσ π( )SCP

dΩ≈

Kc48β( ′ G )2 (7)

c is the speed of light,

β( ′ G )2 is the anisotropic ROA invariant , and K is:

K =107

9π 2µo

2c 4 ˜ ν oωo

200πc−Δ ˜ ν p

3

(8)

Page 18: ABSTRACT - Duke University

18

ωo = 2πν o is the frequency of the incident light,

Δ ˜ ν p is the Raman frequency shift (in cm-1), c is

the speed of light (in m s-1), and

µo is the permittivity of vacuum.

Placzek polarizability theory writes 19

β( ′ G )2 =12

3αµν ′ G µν −αµµ ′ G νν( )µν

∑ (9)

where

αµν ′ G µν is

αµν ′ G µν =h

8π 2cΔ ˜ ν p i, j∑

∂αµνe

∂xiα

0α ,β∑

∂ ′ G µνe

∂x jβ

0

Lαi,px Lβj ,p

x (10)

Lαi,px is the ith component of the Cartesian displacement vector of nucleus α for normal mode p.

αµνe and

′ G µνe are the elements of the electronic components of the electric dipole-electric dipole

tensor and the imaginary part of the electric dipole-magnetic dipole tensor, respectively. The

index 0 indicates that the derivatives of the electronic tensors are evaluated at the equilibrium

nuclear geometry.

In order to explore how G′′ tensor is used to address stereochemical puzzles, we have

divided the following section into two topics: (1) The configurational and conformational analysis

of molecules using chiroptical spectroscopy, and (2) The chiroptical response of molecular

assemblies. These sections also include an overview of the computer simulations of the solvent

effects on OR and ROA, which is the focus of our research, and the details of these simulations

are described in chapters 2, 3 and 4.

Page 19: ABSTRACT - Duke University

19

1.4 Utilizing the G′(ω) tensor to probe molecular dissymmetry

1.4.1 Configuration and conformation analysis of molecules usingchiroptical spectroscopy.

The applications of OR, ECD, and ROA to molecular structure analysis requires the

accurate and efficient computation of chiroptical response properties. We now describe how the

prediction of chiroptical properties in concert with experiments is a productive approach to probe

and to establish links between molecular stereochemistry and chiroptical response.

Determining the absolute configuration (AC) of a chiral molecule is often a considerable

challenge. Common tools to determine the AC of natural products include X-ray crystallography,

ECD 20, nuclear magnetic resonance (NMR) analysis of Mosher ester and relative derivates 21,

chemical degradation and total synthesis 22,23. Chemical degradation and total synthesis are time

consuming, costly, and error prone 24. Comparison of calculated and experimental OR data, in

contrast, provides a straightforward and less expensive strategy to assign AC. Kondru et al.

reported the first ab initio theoretical AC assignment of a natural product (Hennoxazole A) by

computing the molar rotation of its fragments and using van’t Hoff’s superposition principle to

combine the fragment results 25. This approach divided the flexible natural product into weakly

interacting fragments, computed the Boltzmann-averaged molar OR for each fragment, and

summed the contributions to obtain a composite molecular value for comparison with experiment.

Prior experimental studies had validated the use of the additivity principle for this natural product26. This approach was also successfully applied to determine the AC of the natural product

Pitiamide A 27. Recently, Zuber et al. combined NMR NOEs, chemical shift, and J-coupling

measurements with both OR and ECD data to determine the AC of the marine natural product

Bistramide C. Bistramide C has 10 stereogenic carbons with, 1024 possible stereoisomers 28. The

Bistramide C study demonstrates that the judicious use of electronic structure analysis and

spectroscopic data can be used to accomplish successful AC assignment of complex flexible

natural products with a large numbers of stereogenic carbons. Other groups have also noted the

Page 20: ABSTRACT - Duke University

20

utility of using multiple chiroptical methods, including ORD, ECD, VCD, and ROA to determine

the AC of chiral species 29.

The reliability of computational methods for predicting OR was demonstrated by assigning

the yet undetermined stereochemistry of several molecules. For example, Zuber et al. used TD-

DFT, OR calculations to assign (+) and (-) OR signs to trans-(S)- and trans-(R)-ladderane

configurations, respectively 30. Recently, Mascitti and Corey carried out the enantioselective

synthesis of pentacycloanammoxic acid with a trans-(S)-ladderane constituent 31, which was

indeed found to have a positive [α]D, consistent with our prediction.

The development of computational methods to describe chiroptical response is enabling the

examination of molecular conformations and assemblies via these responses. Kondru, Beratan

and Wipf defined atomic and chemical group contributions to OR to establish an intuitive link

between structure and OR 32. Atomic contribution maps indicate that OR calculations are

sensitive to molecular geometry, and thus, reliable prediction of molar rotation for flexible

molecules requires extensive conformational averaging 33. Indeed, Vaccaro and co-workers find a

strong dependence of computed OR on molecular conformations in 2-substituted butanes 34,

similar to earlier analysis of OR in twisted “molecular wires” 35.

The availability of commercial ROA spectrometer and corresponding computational methods

has made ROA spectroscopy a useful tool for configuration and conformation analysis of chiral

molecules 14,17. We have demonstrated the utility of ROA by predicting the distribution of

dominant N-acetylalanine-N′-methylamide (Ala dipeptide) conformations in aqueous solution 36

(see chapter four for more details).

The conformational heterogeneity of Ala dipeptide has been probed using multiple

spectroscopic methods in addition to ROA. Despite extensive study, there is no consensus for Ala

dipeptide’s structure in aqueous solution. For example, the 13C NMR spectra of Ala dipeptide in

water and in liquid-crystalline media suggest that the structure in these two media is similar, and

that the structure is dominated by extended left-handed PPII helical conformations 37,38. Following

these NMR studies, Mehta et al. used 13C NMR 39 to show that Ala dipeptide in water exists in a

Page 21: ABSTRACT - Duke University

21

mixture of PPII and the compact right-handed αR helical conformations. Kim et al. probed the

conformational heterogeneity of Ala dipeptide in aqueous solution using 2D-IR 40 and reported

the presence of PPII-like conformations, and ruling out αR helical conformations in water. Raman

spectroscopy, in contrast, suggests that PPII, αR and the intra-molecular hydrogen bonded C7eq

conformations coexist in aqueous solution 41.

The significant variation of proposed Ala dipeptide conformations warrants further analysis,

and suggests the utility of linking chiroptical simulations to experimental data. ROA has been

used to characterize aqueous solution conformations of the peptide backbone in polypeptides and

proteins 42. The ROA spectrum reports a weighted superposition of spectra arising from molecular

conformations 42. We analyzed the ROA spectra of Ala dipeptide in H2O and D2O using TD-DFT

on Monte Carlo (MC) sampled geometries. We thus explore the consistency of MC generated

structures with experimental ROA data. Our analysis of the Raman and ROA spectra of Ala

dipeptide indicate that it is dominated by αR and PPII conformations in aqueous solution 36.

1.4.2 Chiroptical signatures of intra- and inter-molecular interactions.

Atomic and chemical group contributions to OR indicate that functional groups far from

stereogenic centers can dominate the OR of chiral species. For example, Kondru et al. found that

the large difference in the OR of the marine natural products Calyculin A and B, which differs

only by the (E)- vs (Z)-configuration of a terminal double bond in a tetraene unit distant from the

stereogenic center, arises from the polarizability of the terminal cyanide (CN) substituent rather

than by direct perturbation of the stereogenic carbons 43. The Calyculin case demonstrates that π-

mediated intramolecular electronic communication between the stereocenter and a distant CN

group dominates the OR.

A key challenge in stereochemical analysis is to compute chiroptical responses accurately

for molecules in solutions. Indeed, OR in solution is known in many instances to be very different

from the response in the gas phase 44. Condensed phase effects arise from 1) the chiral solvent

imprint, 2) the perturbation of the chiral species by solvent, 3) molecular assembly, and 4) the

Page 22: ABSTRACT - Duke University

22

influence of solvation on the population of molecular conformers and concominant changes in

electronic structures.

Our studies have shown that the solvent effects on ORD can be modeled using a

combination of structure sampling and electronic structure methods (see chapter two and three for

model details). We demonstrated the validity of this approach in studies of the solvent

dependence of ORD for methyloxirane, a particularly difficult and relevant case that has

confounds some of the highest-level computational methods. (S)-methyloxirane has a positive

ORD spectrum in water and a negative spectrum in benzene in the long wavelength regime 45. A

hydrogen bonded methyloxirane-water adduct dominates the ORD in water 46. For methyloxirane

in benzene, the chiral imprint on the benzene cluster dominates the ORD spectrum 47. That is,

removing the methyloxirane and computing the ORD of the imprinted benzene snapshots

reproduces the observed ORD spectra.

Xu and coworkers have recently studied the solvation of methyloxirane in water using

VCD, OR, and simulation 48. They concluded that the methyloxirane-water binary complex

dominates the chiroptical signature in aqueous solution at room temperature, and the anti

conformation is preferred over the syn conformation, consistent with our predictions 46. The OR

of (R)-pantolactone is also controlled by noncovalent association via hydrogen bonding. Indeed,

(R)-pantolactone in carbon tetrachloride at room temperature was shown to have [α]D that arises

from a combination of monomers and dimers 49.

The ORD signature of methyloxirane in benzene arises from a chiral imprint on the

solvent, not from perturbation of methyloxirane by the solvent 47. The methyloxirane-benzene

case demonstrates that a chiral solvent cluster seeded by a chiral solute may dominate the

chiroptical response. Indeed, chiral solutes are known to induce chiral solvent ordering that

contributes to the chiroptical response. Fidler et al. showed that dissymmetric solvent

organization around a chiral solute accounts for 10-20% of the total CD intensity attributable to

an optically active chromophore 50. Yet, the dominance of an induced chirality signature was

shown for the first time using the theoretical computation of methyloxirane’s ORD in benzene 51.

Wang and Cann have recently studied the extent of induced chirality in achiral solvent by chiral

Page 23: ABSTRACT - Duke University

23

solute using MD simulations 52. They concluded that chirality transfer is most pronounced for

solvents that are hydrogen bonded to solutes, and the dissymmetric solvent shell around the solute

at specific positions also contributes to the extent of chirality transfer.

The chirality of metals or nanocluster with chiral adsorbates is the subject of intense

interest because of potential applications in enantioselective catalysis 53, enantiodiscrimination 54,

enantioselective crystallization 55 and photonics 56. Theoretical analysis indicates that the chiroptic

response of metal clusters with chiral adsorbates may arise from an intrinsically chiral core

structure of the metal cluster 57, a chiral surface or chiral adsorbates. Goldsmith et al. employed

simple model calculations using a dissymmetrically-perturbed particle-in-a-box model to probe

chiral imprinting of nanoparticles. The authors found that CD may be induced in a nanoparticle

by a chiral monolayer as in Au28(SG)16 glutathione passivated gold nanoclusters 58. Similar to

solution chiral imprints, the analysis of the CD spectra suggests that the chiroptical response of

the chiral monolayer protected cluster can arise from the imprinted core lattice with dissymmetric

adsorbates. Gautier and Bürgi recently showed, using thiolate-for-thiolate ligand exchange

experiments on gold nanoparticles, that the optical activity of gold nanoparticles is indeed

dictated by the chirality of the adsorbed thiolates 59, consistent with theoretical predictions 58.

1.5 Summary

Three-dimensional shape is a key determinant of molecular function. Theoretical

calculations of OR, ORD, ECD, and ROA, when combined with experiments, can be used to

assign stereochemistry (both conformation and configuration) for natural products and

biomolecules.

A key challenge in stereochemical analysis is to compute the chiroptical response of

molecules in solution, which is the focus of our research in optical activity. Our theoretical

studies of methyloxirane in water and benzene, which are described in chapters 2 and 3, show that

hydrogen bond and dispersion interactions strongly influence the chiroptical responses 46,47.

Although the common perception is that chiroptical responses are solely determined by a chiral

Page 24: ABSTRACT - Duke University

24

solute’s electronic structure in its environment, we demonstrate that chiral imprinting effects on

media, and molecular assembly effects can dominate chiroptical signatures. Both these effects are

strongly influenced by intermolecular interactions that are essential to accurately describe

chiroptical signatures of molecules. Chiroptical response of molecules spans a range of large

positive and negative values 46,47,60, and hence, the measured chiroptical response represents an

ensemble average of orientations. Thus, accurate prediction of chiroptical properties requires

adequate conformational averaging. Methods for computing chiroptical response in the

condensed phase using continuum solvation models and point charge solvent model, although

more efficient than explicit solvent treatments, are often inadequate for describing induced

chirality effects 47. The lack of a priori knowledge of solvent-solute interactions and their

influence on the chiroptical signature demands exploration of thermally averaged solute-solvent

clusters and comparison with simpler solvent studies.

Page 25: ABSTRACT - Duke University

25

2 Solvent effect on optical rotation: A case study ofMethyloxirane in water

2.1 Overview

In order to understand solvent effects on optical rotation (OR), the solvent dependence of

the ORD for methyloxirane was studied. The figure above shows the measured (dashed line) and

computed (solid line) ORD of (S)-methyloxirane in water () and benzene (). The hydrogen

bonded methyloxirane-water adduct (top structure) dominates the ORD in aqueous solution. In

contrast, the chiral benzene cluster (bottom structure) dominates the ORD in benzene (also see

chapter 3 for more details). Here we show that including explicit solute-solvent interactions is

essential to describe the influence of the solvent on the OR of methyloxirane in water.

2.2 Introduction

A long-standing challenge in molecular stereochemistry is to assess the contributions of

the solvent surrounding chiral organic molecules. The observed specific rotation angles and ORD

are strongly influenced by solvent-solute interactions. For example, (S)-methyloxirane has a

positive OR in water and a negative OR in benzene 45. The vapor phase and condensed phase

Page 26: ABSTRACT - Duke University

26

optical rotations of chiral molecules are also known to differ in large and nonintuitive ways 61.

Access to experimental gas phase optical rotation data 44 and to modern linear-response OR

calculations 11,13,25,33,34,49,62-67 provides the tools needed to analyze and predict the contributions of

solute-solvent interactions to OR. Wilson et al. recently reported the OR for a series of organic

molecules in the vapor phase using cavity ring-down polarimetry 44. These authors compared the

solution and vapor phase OR to assess the influence of the condensed phase environment on

ORD. They concluded: (a) for flexible solute molecules, solvent stabilization of individual

conformers can influence the optical rotation strongly; (b) solvation of rigid molecules like

methyloxirane causes changes in the static and dynamic distribution of molecular electron

densities; and (c) surprisingly, OR under solvent-free conditions can resemble OR measured in a

highly polar solvent. We find that, for methyloxirane in high-polarity solvents such as water, it is

essential to include explicit solute-solvent interactions in the theoretical analysis to describe the

observed ORD.

Recently, coupled-cluster (CC) and time dependent density functional theory (TD-DFT)

methods have been used to predict OR 9. However, the substantially greater computational

efficiency of TD-DFT methods has led to their favored use in current OR calculations. The

solvent dependence of OR for chiral molecules has been computed using TD-DFT and an implicit

continuum solvent model. Kongested et al. recently used CC methods combined with a

continuum solvent description to calculate the influence of solvent on the OR of (S)-

methyloxirane 68. The authors concluded that continuum models do not reproduce the

experimentally observed solvent shifts in the ORD as a function of solvent. Moreover, implicit

solvent models do not predict the sign change in the ORD of methyloxirane in aqueous solution at

wavelengths longer than 400 nm. Accordingly, the inclusion of explicit solvent effects appears to

be essential for further progress. Here, we show that ORD calculations using explicit solvent

models, combined with continuum models, such as the conductor-like screening model

(COSMO) 69, can capture the sign and relative magnitude of the ORD of methyloxirane in water.

More importantly, explicit solvent studies allow elucidation of the solvent contributions to the

observed OR spectrum. Comparison of the total computed OR (solute + solvent) with the OR

Page 27: ABSTRACT - Duke University

27

computed for the solvent shell (solvent with the chiral solute removed from the electronic

structure analysis) allows us to determine when explicit solute-solvent interactions have a critical

effect on the ORD.

Ruud and Zanasi 70 have shown that it is important to include molecular vibrational

effects to correctly predict the experimental gas phase OR values of (S)-methyloxirane

successfully. Ruud et al, however, showed that vibrational corrections in dimethyloxirane

decrease the accuracy of the calculated OR values, which was attributed to the absence of solvent

effects 66. More recently, Mort and Autschbach 67 showed that including vibrational corrections

results in a small percentile difference between the experimental solution phase numbers and the

calculated gas phase values with or without the vibrational effects in a series of rigid molecules.

They attributed this behavior to the absence of solvent effects in their OR calculations. Rather

than making vibrational corrections to our OR calculations, we focused entirely on the influence

of solvent effects. As discussed by Mort and Autschbach, OR calculations incorporating both

vibrational and solvent effects might result in a further improved quantitative agreement between

theory and experiment.

Page 28: ABSTRACT - Duke University

28

2.3 Results and Discussion

Figure 2.1 Optical rotatory dispersion of methyloxirane in the gas phase and inwater computed using an implicit solvent model (A) and a combination of implicit andexplicit solvent models (B). Calculations used the TD-DFT/BP86 functional and aug-cc-pVDZ basis set. The conductor-like screening model (COSMO) was used as the implicitsolvent model. The vertical bars show the error estimate of the average optical rotation

calculated using the blocking method 71.

Page 29: ABSTRACT - Duke University

29

Figure 2.1 shows the computed ORD of (S)- and (R)-methyloxirane in water and in the

gas phase. The ORD were calculated using the BP86 functional with an aug-cc-pVDZ basis set.

Each structure had explicit water molecules within a cut-off distance of 0.35 nm from the

methyloxirane oxygen atom and a dielectric continuum for large distances based on the COSMO

model (water dielectric constant equal to 78.4). The computed ORD of (S)- and (R)-

methyloxirane in water were related by a simple reversal of sign, as expected. The ORD

predictions with explicit solvent were very close to the experimental values, especially for

wavelengths >400 nm. Calculations using the implicit solvent model only (Figure 2.1 A) predict

the wrong sign for the OR of methyloxirane at 400 nm and longer wavelengths, while

calculations using a combination of explicit and implicit solvent models (Figure 2.1 B) correctly

predict the sign of the OR at all wavelengths. Interestingly, in the water-methyloxirane system

(Figure 2.1 B), the OR of the pure solvent shell (water + COSMO without the chiral solute) is a

minor fraction of the total OR system (solute + water + COSMO) at all wavelengths. Thus, the

water-methyloxirane interactions in an aqueous methyloxirane solution constitute the major

contribution to the observed ORD.

Solvent significantly influences the OR of chiral molecules. For example, Gottarelli et al.

showed that chiral biaryls show significant variation in OR in biphenyl-type solvents 72. Based on

a statistical theory of dielectric susceptibility and optical activity of isotropic liquids the authors

described three different solvent effects on OR. First, the solvent manifests itself as a continuous

dielectric medium characterized by its refractive index. Second, the induction of chiral

conformations in the neighbouring solvent molecules contributes to the OR. And third, short-

range interactions between solute and solvent influence the OR in solution. Craig and

Thirunamachandran showed that intermolecular interactions between chiral and achiral molecules

results in the modification of the optical rotatory tensor of the chiral molecule, such that it

displays changed chiroptical properties 73. Here we show that the optical response of

methyloxirane in water derives from hydrogen bonded methyloxirane-water adduct and not from

methyloxirane alone. Goldsmith et. al. showed that in order to compute the correct optical

rotation of pantolactone, it is important to compute the optical rotation of the pantolactone dimer,

Page 30: ABSTRACT - Duke University

30

which exists in equilibrium with the monomer in solution, and the response originates from the

dimer adduct and not just the individual monomer 49. Thus, we emphasize that it is important to

include explicit solvent models in order to describe solvation effects on optical rotation properly,

because the optical response may arise from the solute-solvent adduct (as in the case of aqueous

solution methyloxirane), which is a very different chemical species than the chiral solute alone.

Figure 2.2 Radial distribution function of hydrogen atoms for water around themethyloxirane oxygen atom.

Figure 2.2 shows the calculated radial distribution function of the hydrogen atoms of

water molecules around the methyloxirane oxygen atom. This function shows a well-defined

minimum at 0.25 nm. We included explicit solvent molecules within a cut-off distance of 0.25 nm

from the oxygen atom of methyloxirane to model the first solvation shell. Subsequently, more

water molecules were added to our calculations with cut-off distances of 0.35 and 0.40 nm. Thus,

we also computed OR values with different numbers of water molecules in the quantum

mechanical analysis. The total number of water molecules within the cut-off distances was 1-2

(0.25 nm), 3-5 (0.35 nm) and 6-9 (0.40 nm).

Page 31: ABSTRACT - Duke University

31

Figure 2.3 shows the variations of the computed OR of (S)-methyloxirane as a function of

the number of solvent molecules. These ORD spectra were calculated using the BP86 functional

and an aug-cc-pVDZ basis set. Apart from the OR at 365 nm, we found no significant variation in

the computed ORD of the aqueous methyloxirane solution with the number of solvent molecules

included. The OR at 365 nm approaches the experimental value as more water molecules are

included in the calculation up to the maximum cut-off distance of 0.40 nm. We also emphasize

that the computed average OR with explicit water molecules within a cut-off distance of 0.40 nm

from the methyloxirane oxygen atom, with and without the implicit solvent model (COSMO) at

589 nm, [α]D , was 8.3 and 7.3, respectively.

Figure 2.3 Optical rotatory dispersion of (S)-methyloxirane in water computed withan increasing number of solvent molecules. Calculations used the TD-DFT/BP86 functionaland aug-cc-pVDZ basis sets. The conductor-like screening model (COSMO) was used as the

implicit solvent model. The error bars were calculated using the blocking method 71.

Previous theoretical studies suggested that TD-DFT methods with the B3LYP functional

and (at least) an aug-cc-pVDZ basis set are required to make reliable OR predictions 74. Thus, we

also compared the results of our OR calculations with different combinations of BP86/B3LYP

correlation-exchange functionals and aug-cc-pVDZ /aug-cc-pVTZ basis sets. We included only

the first solvation shell around methyloxirane in the comparison, as we found no significant

Page 32: ABSTRACT - Duke University

32

change (except at 365 nm) in the computed ORD as a function of the number of solvent

molecules included (see Figure 2.3).

Figure 2.4 Variation in the dipole-length and dipole-velocity form of calculatedoptical rotations using TD-DFT/BP86 correlation-exchange functional. The specific rotation

was averaged for 1700 structure snapshots from a molecular dynamic simulation togenerate the ORD of (S)-methyloxirane in water. The conductor-like screening model

(COSMO) was used as the implicit solvent model. The vertical bars show the error estimateof the average optical rotation calculated using the blocking method 71.

Page 33: ABSTRACT - Duke University

33

Figure 2.4 shows the computed ORD of (S)-methyloxirane with the first hydration shell,

which includes explicit solvent molecules within a cut-off distance of 0.25 nm from

methyloxirane oxygen atom. The dipole-length (L) and dipole-velocity (V) forms (see Method

section for details) of the TD-DFT results with BP86/aug-cc-pVDZ (or aug-cc-pVTZ) are very

similar. Thus we expect the L and V forms of the results with B3LYP/aug-cc-pVDZ (or aug-cc-

pVTZ) will also be very similar.

Figure 2.5 Computed optical rotatory dispersion of (S)-methyloxirane with the firstexplicit hydration shell. Calculations used different combinations of TD-DFT/BP86/B3LYP

functionals and aug-cc-pVDZ/aug-cc-pVTZ basis sets as described in the text. Theconductor-like screening model (COSMO) was used as the implicit solvent model. The bars

indicate error estimates computed using the blocking method 71.

Figure 2.5 shows that the computed ORD spectra of (S)-methyloxirane with the

BP86/aug-cc-pVDZ (or aug-cc-pVTZ) and B3LYP/aug-cc-pVDZ (or aug-cc-pVTZ) analysis are

very similar, especially at wavelengths longer than 400 nm We emphasize that the L form of TD-

DFT, as implemented in Turbomole5.6 with the B3LYP exchange-correlation functional, was

used to compute the OR. However, the L and V forms of the TD-DFT results with BP86/aug-cc-

pVDZ (or aug-cc-pVTZ) are very similar (see Figure 2.4). Thus, we expect the computed L- and

Page 34: ABSTRACT - Duke University

34

V- based values with B3LYP/aug-cc-pVDZ (or aug-cc-pVTZ) also to be very similar. These

results show that the computed ORD of (S)-methyloxirane with the BP86/aug-cc-pVDZ (or aug-

cc-pVTZ) and B3LYP/aug-cc-pVDZ (or aug-cc-pVTZ) analysis are very similar, especially at

wavelengths longer than 400 nm.

Figure 2.6 Average optical rotation of (S)-methyloxirane at 589 nm, [α]D, as afunction of the number of MD simulation snapshots. Calculations used the TD-DFT/BP86functional and aug-cc-pVDZ basis set. A combination of explicit water molecules (within

0.40 nm of the methyloxirane oxygen atoms) and an implicit solvent model (COSMO) wereused in the calculations. The error estimate of the average optical rotation was calculated

using the blocking method 71.

Figure 2.6 shows the average computed [α]D value as a function of the number of MD

simulation snapshots. Each structure had explicit water molecules within a cut-off distance of

0.40 nm from methyloxirane oxygen atom with a dielectric continuum description based on the

COSMO model with the water dielectric constant of 78.4. The average [α]D value converged to

+8.4 as the number of snapshots were increased.

Page 35: ABSTRACT - Duke University

35

Figure 2.7 Distribution of the optical rotation of (S)-methyloxirane at 589 nm, [α]D,as a function of water molecule positions defined by the angle τ, in the first hydration shell.

The insets show structures with water molecules on the same or opposite sides from themethyl group of methyloxirane, defined as the syn- or anti-configurations, respectively.

Water on the opposite side dominated the explicit solvent contribution to [α]D.

In order to understand the physical origin of the positive OR for methyloxirane in water,

we computed the OR at 589 nm and displayed [α]D as a function of the angular position of the

water molecules, defined by the angle τ (see inset in Figure 2.7), in the first hydration shell (there

is only a minor dependence of the computed [α]D on the number of solvent molecules, see Figure

2.3). Structures with water molecules adjacent to or opposite to the methyl group of

methyloxirane are defined as syn- or anti-configurations, respectively (see insets in Figure 2.7).

Figure 2.7 shows the τ histogram for 14,000 MD structure-snapshots. The [α]D values were

calculated with the BP86 functional and the aug-cc-pVTZ basis set. The distribution shows that

the syn-configurations make a nearly equal number of positive and negative [α]D contributions,

while the anti-configurations make a predominantly positive contribution to [α]D contributions.

Based on the MD simulations, therefore, the positive [α]D value for methyloxirane in water arises

Page 36: ABSTRACT - Duke University

36

from the greater number of water anti-configurations with positive [α]D. It is also important to

note that the average [α]D values for 14,000 and 1,700 structure snapshots are +10.4 and +9.8,

respectively.

Su and Xu recently showed that for methyloxirane-(water)2 ternary cluster in the gas

phase, the anti conformer is more populated than the syn conformer 75. In contrast, the syn

conformer is more populated for the methyloxirane-water binary cluster in the gas phase 76. The

authors suggested that the secondary hydrogen-bonding of the methyl group and water, which

results in the syn conformational preference for methyloxirane-water binary cluster, do not

contribute to the stability of the methyloxirane-(water)2 ternary cluster. For the ternary cluster,

maximizing the interactions between the second water and methyloxirane results in the

conformational preference of anti conformer, which is consistent with our theoretical prediction.

Following our OR calculation of methyloxirane aqueous solution, Xu and coworkers also studied

the solvation of methyloxirane in water using VCD, OR and computer simulations 48. Those

authors concluded that the methyloxirane-water binary complex is the dominating species in

aqueous solution at room temperature, which is also consistent with our prediction 46.

2.4 Conclusions

Molecular dynamics simulations and time-dependent density functional theory (TD-DFT)

methods were used to determine the solvent dependence of the optical rotatory dispersion (ORD)

spectrum for methyloxirane in water. The off-resonance ORD spectrum was calculated using both

explicit and conductor-like screening solvent models (COSMO). TD-DFT ORD calculations at

the BP86/aug-cc-pVDZ (or aug-cc-pVTZ) and B3LYP/aug-cc-pVDZ (or aug-cc-pVTZ) levels

using explicit water molecules reveal excellent agreement between experiment and theory in both

sign and magnitude; the COSMO model alone does not adequately describe the influence of

solvent on the ORD spectrum of methyloxirane. The positive OR at 589 nm, i.e. the [α]D value, is

attributed to the greater number of MD snapshots with water molecules on the side of the

methyloxirane ring opposite to the methyl group. We conclude that including explicit solute-

solvent interactions is essential to describe the influence of the solvent on the OR of

Page 37: ABSTRACT - Duke University

37

methyloxirane in water and that MD simulations provide a suitable means to analyze and predict

chiroptical solvation effects.

2.5 Materials and methods

2.5.1 Experimental Procedure

Optical rotation (α) was measured with an AUTOPOL IV digital polarimeter. Optical

rotation of (S)-methyloxirane (Alfa Aesar) in water and benzene at 0.05 g/mL was measured at

six different wavelengths. The specific rotation at a given wavelength [α]λ was calculated using:

α[ ]λ =αVml

α is the optical rotation in degrees, V is the volume (ml) containing a mass m (g) of the

optically active substance, and l is the path length (dm).

2.5.2 Molecular dynamics simulation setup and procedure

Molecular dynamics (MD) simulations of (S) and (R)-methyloxirane oxide in a pre-

equilibrated box of 2175 water molecules were done in a NPT ensemble using Gromacs 3.1.4. 77

In all simulations a leapfrog integrator was used with a 2-fs time step 78. All bonds were

constrained to their equilibrium value with the SETTLE algorithm 79 for water and the LINCS

algorithm 80 for all other bonds. In all simulations, a DFT/B3LYP/6-311G** optimized fixed

geometry of (S)-methyloxirane was used. A twin-range cutoff was used for the Lennard-Jones

interactions, with interactions within 0.9 nm evaluated every step and interactions between 0.9

and 1.4 nm evaluated every 5 steps. A neighbor list was used and updated every five steps. For

electrostatic interactions, a reaction field with a Coulomb cutoff of 1.4 nm was used with

dielectric constants of 78.4 for water beyond the cutoff. The pressure was held at 1 bar using the

weak-coupling scheme with a coupling constant of 1.0 ps and an isothermal compressibility of

4.5 ×10-5 bar-1. Each component of the system (i.e. (S)-methyloxirane and water) was coupled

separately to a temperature bath at 300 K, using the Berendsen thermostat 81, with a coupling

Page 38: ABSTRACT - Duke University

38

constant of 0.1 ps. The SPC water model was used 82. The OPLS-AA force field 83 was used and

partial charges on (S)-methyloxirane were calculated in Gaussian03 at the DFT/B3LYP level with

the 6-31G* basis set and charge partitioning using the chelpg method 84. Simulations in water

were carried out for 40 ns.

2.5.3 Computation of optical rotation

Structures from the MD trajectories were used for optical rotation (OR) calculations. The

specific rotation was averaged for 1700 structure snapshots at each wavelength to generate the

ORD spectrum of methyloxirane in water. Specific rotations for each structure were calculated

according to the equation:

α[ ]λ =−1.34229 ×10−4Tr ′ G ω( )[ ]ν 2

3Mω

where Tr [G (ω)] is the trace of frequency dependent electric-dipole magnetic-dipole

polarizability tensor,

ν (=λ-1=ω/2πc) is the radiation wavenumber, and M is the combined molar

mass of (S)-methyloxirane and the number of solvent molecules in each snapshot. Thus, by

including the mass of the water molecules we are properly normalizing [α]λ with respect to the

total number of water molecues. [α]λ, is an intensive property independent of the number of

solvent molecules. The dipole-length (L) and dipole-velocity (V) formulation of time dependent

density functional response theory (TD-DFT) as implemented in Turbomole5.6 85 with B3LYP

and BP86 exchange-correlation functional, respectively, was used for the calculation of [G’ (ω)].

Page 39: ABSTRACT - Duke University

39

3 Contribution of a solute’s chiral solvent imprint to opticalrotation

3.1 Overview

Solvent or solute dissymmetry? Dissymmetric solvent ordering around a chiral molecule

(see Figure) in solution contributes to the chiroptical signature. Indeed, the solvent can dominate

the chiroptical response, as shown for (S)-methyloxirane in benzene.

3.2 Introduction

A long-standing challenge in molecular stereochemistry is to assess the “chiral imprint”

of a chiral solute on the surrounding solvent. The solute’s influence on ordering the solvation

sphere should contribute to the optical rotatory dispersion (ORD). The magnitude of this

contribution, however, has never been assessed. We now show that for (S)-methyloxirane in

benzene, the chiral imprint in the solvent sphere dominates the optical rotation (OR). To the best

of our knowledge, this is the first evidence in support of a chiroptical property dominated by the

dissymmetry induced in the solvent.

The observed specific OR angles and ORD of chiral molecules in solution are well

known to be strongly influenced by solvent-solute interactions. For example, (S)-methyloxirane

has a positive OR in water and a negative OR in benzene 45. Thus, theoretical modelling to

understand both the sign and magnitude of OR is of interest. Access to experimental gas and

Page 40: ABSTRACT - Duke University

40

solution phase OR data 61 and to modern linear-response OR calculations 11,13,25,33,34,49,62-67 provide

the tools needed to dissect solute and solvent contributions to OR. In the preceding chapter, we

showed that water-methyloxirane interactions in an aqueous solution dominate the observed ORD46; in contrast, for methyloxirane in benzene, we show here that the chiral solvent ordering is the

dissymmetry that dominates the ORD.

The solvent dependence of OR for chiral molecules has been computed recently using

coupled-cluster (CC) and time dependent density functional theory (TD-DFT) using implicit

continuum solvent methods 9. Kongested et al. used CC combined with a continuum solvent

description to calculate the influence of solvent on the OR of (S)-methyloxirane 68. The authors

concluded that continuum models: (a) do not reproduce the experimentally observed solvent

shifts of the ORD spectra as a function of solvent; and (b) cannot describe the contribution of the

chiral solvent structure induced by the chiral solute to the observed OR. Explicit solvent models

capture solvent/sign anomalies in the OR and allow attribution of contributions to OR arising

from a solute’s chiral imprint on its environment as shown below.

Page 41: ABSTRACT - Duke University

41

3.3 Results and Discussion

Figure 3.1 Optical rotatory dispersion (ORD) of (S) ()- and (R) ()-methyloxirane-benzene clusters (shown in the MC structure snapshot), and the solventimprint () computed using explicit solvent models. Calculations used TD-DFT/BP86

functional and SVP basis set; and an implicit solvent () based on the COSMO model. Theerror bars are calculated using the blocking method 71.

Figure 3.1 shows the computed ORD spectra of (S)- and (R)-methyloxirane in benzene

using explicit and implicit solvent models. The ORD spectra were calculated using the BP86

functional with the SVP basis set. Calculations using a dielectric continuum based on the

COSMO model 69 (an implicit solvent model) with a benzene dielectric constant of 2.0 incorrectly

predict the sign of the OR approaching resonance from long wavelengths. Explicit solvent,

however, correctly predicts the sign of the OR at these wavelengths.

The computed ORD spectra of (S)- and (R)-methyloxirane based on explicit solvent

models are related by a simple reversal of sign, as expected. Interestingly, in the methyloxirane-

benzene system, the OR of the solvent imprint (benzene cluster without the chiral solute) is

comparable to the total OR of the system (solute + benzene) at all wavelengths (Figure 3.1).

Thus, the chiral solvent ordering of benzene, or the chiral imprint, dominates the ORD. Chiral

solutes are known to induce chiral solvent ordering. Fidler et al., for example, showed that

Page 42: ABSTRACT - Duke University

42

dissymmetric solvent organization around a chiral solute accounts for 10-20% of the total circular

dichroism intensity attributable to an optically active chromophore 50,86. Our ORD calculations for

methyloxirane in benzene attribute the chiroptical signature to the dissymmetric benzene cluster

around the methyloxirane. Morimoto et al. recently showed that the attractive interactions

between benzene rings results in the formation of a cyclic trimeric dissymmetric benzene cluster

in solution, which is reminiscent of the assembles seen in our MC simulations 87.

Computed OR values are very sensitive to molecular conformation 33,34. Thus, the

difference in the computed and experimental OR values (Figure 3.1) might arise from the

approximations in the description of molecular interactions in the MC simulation of

methyloxirane in benzene arising from the force field limitations as well as approximations in the

quantum mechanical calculations.

Page 43: ABSTRACT - Duke University

43

Figure 3.2 Optical rotatory dispersion (ORD) of (S)-methyloxirane in benzene ()and the computed ORD of the chiral solvent imprint (;;;). Calculations used varied

TD-DFT/BP86/BLYP correlation-exchange functionals with RI-J and SV/SVP/aug-cc-pVDZ basis sets. The specific rotation was averaged over 1000 structure snapshots from a

Monte-Carlo simulation to generate the ORD of the chiral solvent structure.

The ORD spectra of the chiral solvent imprint were also computed using different

combinations of BP86/BLYP correlation-exchange functionals and SV/SVP/aug-cc-pVDZ basis

sets. We found no significant variation in the computed ORD spectra with the choice of

functional/basis set. Figure 3.2 shows that the computed ORD spectra of the chiral solvent

imprint with varied correlation-exchange functionals and basis sets are very similar. Previous

theoretical studies suggested that TD-DFT methods with the B3LYP functional and (at least) an

aug-cc-pVDZ basis set are required to produce reliable gas phase OR predictions 74. We

emphasize that using a non-hybrid DFT functional, such as BP86/BLYP, permits use of the RI-J

approximation implemented in Turbomole5.6 to reduce the computational time by six orders of

magnitude compared to calculations with hybrid functionals such as B3LYP; this is especially

important for performing 20,000 OR calculations of benzene-methyloxirane clusters (with 106-

130 atoms) using varied functionals and basis sets. Ensemble averaged calculation of OR for

chiral molecules in solution gives similar results with pure and hybrid functionals, as previously

Page 44: ABSTRACT - Duke University

44

shown for ORD calculations of methyloxirane in aqueous solution 46. Recently, Hassey et al. used

single-molecule spectroscopy to show that the chiroptical response of molecules spans a range of

large positive and negative values, and hence suggested that in the solution phase, the measured

chiroptical response represents an ensemble average of orientations and solvent interactions 60.

Figure 3.3 Optical rotatory dispersion of ethylene oxide in benzene computed usingexplicit solvent model. Calculations used TD-DFT/BP86 correlation-exchange functionals

with RI-J and aug-cc-pVDZ basis set. The specific rotation was averaged over 1000structure snapshots from a Monte-Carlo simulation to generate the ORD of ethylene oxide

and (S)-methyloxirane () in benzene.

In order to show that the predicted chiral solvent structure in methyloxirane-benzene

solution is not an artifact of the methodology, we exemplify it by correctly predicting an average

OR value close to zero for an achiral ethylene oxide-benzene solution. Figure 3.3 shows the OR

of an achiral ethylene oxide-benzene solution. The ORD was calculated using MC simulation and

TD-DFT (with BP86/aug-cc-pVDZ), as described above for the methyloxirane-benzene system.

Each ethylene oxide-benzene structure snapshot from the MC simulation is dissymmetric, with a

nonzero contribution to the OR: the ensemble averaged OR value, however, is close to zero, as

expected.

Page 45: ABSTRACT - Duke University

45

Figure 3.4 Optical rotatory dispersion of (S)-methyloxirane in alkylbenzenes.

In order to probe methyloxirane’s chiral solvent imprint in solutions of benzene

derivatives, we measured the ORD of (S)-methyloxirane in alklybenzenes (Figure 3.4). Figure 3.4

shows that in methyloxirane-alkylbenzene systems, the magnitude of OR in solution decreases as:

Benzene > Toluene > ortho-Xylene > meta-Xylene > 1,2,4 trimethylbenzene > para-Xylene.

Thus, increasing the number of alkyl groups in benzene decreases the OR value of methyloxirane

in alkylbenzenes. The intermolecular steric hindrance between the methyl groups of

alkylbenzenes could decrease the number of solvents in the chiral solvent structure around

methyloxirane, and hence decrease the chiroptical response originating from the chiral solvent

imprint.

3.4 Conclusions

Our results indicate that the contribution to OR of a dissymmetric solvent imprint can

exceed the OR contribution of the solute itself. Implicit solvent models are not sufficient to

describe the imprint effects because they lack explicit inclusion of the solvent electronic structure.

A judicious choice of solvent modelling is essential to describe the chiroptical signature of

molecules in solution.

Page 46: ABSTRACT - Duke University

46

3.5 Materials and Methods

In our study, Monte Carlo (MC) simulation of solute and solvent combined with TD-DFT

methods were used to compute the OR of (S)- and (R)-methyloxirane. Thus, the off-resonance

ORD of methyloxirane in benzene was calculated using an explicit solvent model. MC

simulations of (S)- and (R)-methyloxirane in an equilibrated box of benzene were performed in a

NPT ensemble, using BOSS 88. The all-atom OPLS-AA force-field 83was used in the MC

simulations. OR calculations were performed using TD-DFT implemented in Turbomole5.6 85

with different combinations of the BP86/BLYP correlation-exchange functionals and

SV/SVP/aug-cc-pVDZ basis sets at four wavelengths. All calculations were performed with the

resolution of identity approximation (RI-J) 89. Grimme showed that accurate TD-DFT predictions

of frequency dependent OR for large molecules can be achieved efficiently with the RI-J

approximation 90; structures used in the TD-DFT analysis were taken from the MC simulations.

Each structure had benzene molecules within a cut-off distance of 0.5 nm from the center-of-mass

of methyloxirane. The total number of benzene molecules within the cut-off distance was 8-10.

The specific rotation was averaged over an ensemble of 1,000 structures at each wavelength to

generate the ORD of (S)- and (R)-methyloxirane in benzene. The error estimate of the average

OR was calculated using a renormalization group blocking method 71.

3.5.1 Monte Carlo simulation setup and procedure

Monte Carlo (MC) simulations of (S)- and (R)-methyloxirane in a pre-equilibrated box of

500 benzene molecules were done using BOSS. Simulations were done with a constant number of

molecules at 25° and 1 atm (NPT ensemble) with the Metropolis algorithm. Calculations were

performed with the periodic boundary condition, and long-range electrostatic interactions were

evaluated by the Ewald method. The OPLS-AA force field with flexible solute and solvent

geometries were used in the simulations. AM1 91 based CM1A 92partial charges scaled by 1.14 for

the solute and solvent were used in the MC simulations. A more detailed description of the MC

move and the atomic charge partitioning scheme can be found in 88.

Page 47: ABSTRACT - Duke University

47

3.5.2 Calculation of optical rotation

Structures from the MC trajectories were used for optical rotation (OR) calculations.

Each structure had benzene molecules within a cut-off distance of 0.5 nm from the center-of-mass

of methyloxirane. The total number of benzene molecules within the cut-off distance was 8-10.

The specific rotation was averaged for 1000 structure snapshots at each wavelength to generate

the ORD spectrum of (S)- and (R)-methyloxirane in benzene. Specific rotations for each structure

were calculated according to the equation:

α[ ]λ =−1.34229 ×10−4Tr ′ G ω( )[ ]ν 2

3Mω

where Tr [G (ω)] is the trace of frequency dependent electric-dipole magnetic-dipole

polarizability tensor,

ν (=λ-1=ω/2πc) is the radiation wavenumber, and M is the combined molar

mass of (S)-methyloxirane and the number of solvent molecules in each snapshot. Thus, by

including the mass of the water molecules we are properly normalizing [α]λ with respect to the

total number of water molecues. [α]λ, is an intensive property independent of the number of

solvent molecules. The dipole-length (L) and dipole-velocity (V) formulation of time dependent

density functional response theory (TD-DFT) as implemented in Turbomole5.6, with B3LYP and

BP86 exchange-correlation functional, respectively, was used for the calculation of [G’ (ω)].

Page 48: ABSTRACT - Duke University

48

4 Characterizing aqueous solution conformations of a peptidebackbone using Raman optical activity computation

1.1 Overview

Mounting spectroscopic evidence indicates that alanine predominantly adopts extended

polyproline II (PPII) conformations in short polypeptides. Here we analyze Raman optical

activity (ROA) spectra of N-acetylalanine-N′-methylamide (Ala dipeptide) in H2O and D2O

using density functional theory on Monte Carlo (MC) sampled geometries in order to examine the

propensity of Ala dipeptide to adopt compact right- (αR) and left-handed (αL) helical

conformations. The computed ROA spectra based on MC sampled αR and PPII peptide

conformations contain all the key spectral features found in the measured spectra. However, there

is no significant similarity between the measured and computed ROA spectra based on the αL and

β conformations sampled by the MC methods. This analysis suggests that Ala dipeptide in water

populates αR and PPII conformations but no substantial population of αL or β structures, despite

sampling αL and β structures in our MC simulations. Thus, ROA spectra combined with the

theoretical analysis allow us to determine the dominant populated structures. Including explicit

solute-solvent interactions in the theoretical analysis is essential for the success of this approach.

Page 49: ABSTRACT - Duke University

49

4.2 Introduction

Conformational analysis of alanine containing polypeptides suggests that the polyproline

II (PPII) is the major backbone conformation in aqueous solution. For example, Kallenbach and

coworkers recently showed that alanine containing peptides possess over 90% extended

conformations (i.e. PPII and β structures) 93. The authors, however, did not rule out the possibility

of the presence of compact (αR, αL) conformations of alanine in aqueous solution, and suggested

that such compact structures are present, albeit not dominant. Here, we use Raman optical activity

(ROA) to examine the propensity of Ala dipeptide, which contains a peptide backbone with a

single alanine, to adopt compact structures like the right- (αR) and left-handed (αL) helical

conformations in aqueous solution. We find that Ala dipeptide adopts both the extended PPII and

compact αR conformations, which are the dominant conformations in aqueous solution.

Experimental and theoretical studies of polypeptides 94 suggest that short polypeptides

can adopt secondary structures in aqueous solution. For example, the hepta-alanine polypeptide

Ac-X2A7O2-NH2 (XAO, X and O denote diaminobutyric acid and ornithine, respectively) in

aqueous solution was suggested to be an ensemble of β-turn, β-strand and PPII conformations 95-

97. A combination of Fourier transform infrared (FTIR), vibrational circular dichroism (VCD),

electronic circular dichroism (ECD), Raman, and NMR spectroscopic studies probed shorter

alanine-based cationic peptides (e.g., trialanine, tetraalanine, and H-AAKA-OH). These studies

indicated that these shorter peptides have a higher fraction of PPII than β-strand or αR

conformations 98. ROA studies of a series of A2-A5 peptides suggest that the PPII population

increases with the number of alanines 99. Taken together, these results provide compelling

evidence that alanine containing peptides in water populates PPII conformations. Molecular

dynamics (MD) simulations of alanine containing peptides in water indicate the sampling of αR

motifs, in addition to the PPII and β conformations 100. Graf’s combined MD and NMR studies,

however, show that there is no detectable population of the αR conformation in aqueous solutions

of A3-A7 peptides and that these short polyalanine peptides in water populate the PPII

Page 50: ABSTRACT - Duke University

50

conformation almost exclusively 101. Following Graf’s conformational analysis of A3-A7 peptides,

Hummer and co-worker studied polyalanine peptide (A5) conformations in aqueous solution using

MD simulations in order to address the question of whether the commonly used force fields

overpopulate the α-helical conformations of polyalanine peptides 102. They concluded that the

NMR data are consistent with force fields that give a small α-helical population of polyalanine

peptide and do not require exclusive formation of the PPII structure. Our combined MC and ROA

studies of Ala dipeptide show that the shortest alanine-containing peptide exists as a mixture of

PPII and αR conformations in aqueous solution.

Chiroptical properties are sensitive to molecular geometry and to solvation 33,46,47. ROA is

a chiroptical spectroscopy that measures the difference in the intensity of Raman scattered right-

and left-circularly polarized light, and is utilized to characterize aqueous solution conformations

of the peptide backbone in polypeptides and proteins 42,103. ROA studies of polypeptides suggest

that water promotes structural fluctuations between αR, PPII, and β conformations 104. The ROA

spectrum is the superposition of spectra arising from all conformations of polypeptides in solution42, and thus, ROA analysis can be used to characterize solution conformations of peptide units in

polypeptides.

The theoretical framework for ROA is well known 15,65,105. Accurate time dependent DFT

(TD-DFT) calculations of ROA are accessible using the resolution of identity approximation (RI-

J) 17,90,106 and the BLYP functional with the rDPS basis set 107. ROA spectra are predicted to be

very sensitive to both molecular conformation 14,108-111 and solvation, which were studied

previously using continuum solvent models 112. Herrmann et al. and Pecul et al. analyzed ROA

spectra of polypeptides using a single conformation in the gas phase 113 and in an implicit solvent112, respectively. Since ROA studies of polypeptides indicate that water facilitates conformational

fluctuations 104, modeling polypeptide conformations in explicit solvent, and subsequently

utilizing modelled structures to compute ROA spectra, may improve the effectiveness of ROA

simulations for polypeptides.

Despite extensive experimental and theoretical study, there is no consensus on Ala

dipeptide’s structure in aqueous solution. NMR studies are consistent with both: (1) an equal

Page 51: ABSTRACT - Duke University

51

population of PPII and αR conformations 39,114 and (2) a dominant PPII conformation 37,38. The

later composition is supported by two-dimensional (2D) IR spectroscopy 40. However, Raman

spectroscopy suggests that C7eq, PPII, and αR conformations all exist in aqueous solution 41. MC

and MD simulations of Ala dipeptide indicate either (1) PPII 100,115,116 and αR structures dominate

117-120 or (2) a comparable population of PPII and αR structures dominates 121-123, depending on the

force field used. Thus, a debate remains as to whether PPII is the dominant conformation of Ala

dipeptide in aqueous solution. The significant variation in the description of Ala dipeptide

conformations warrants further analysis, and suggests the utility of linking chiroptical simulations

with experimental studies. Our ROA simulations combined with experimental data show that Ala

dipeptide populates both the αR and PPII conformations in aqueous solution, and hence, it is

erroneous to ascribe all experimental signal to one dominant conformation, namely the PPII.

We present TD-DFT calculations of back-scattered circular polarization (SCP) ROA

based on MC simulation of Ala dipeptide in aqueous solution. Simulated Raman and ROA

spectra constructed using the collection of H2O-Ala clusters (see methods section for more

details) with αR and PPII conformations of Ala dipeptide are consistent with experiments.

However, there is no significant similarity between the measured and computed Raman or ROA

spectra for the set of H2O-Ala clusters with αR and β conformations of Ala dipeptide. Thus, these

calculations facilitate assignment of the observed ROA spectral features to corresponding

secondary structures that characterize the populated conformations of the peptide in aqueous

solution.

Previous ROA analysis using four Ala dipeptide conformations (β, αR, αL, PPII), each

hydrogen bonded to four water molecules and embedded in an implicit solvent model, showed

that the computed spectra of the PPII structure best explained the experimental ROA spectrum 124.

It was also suggested that αR, αL, and “other conformations” of Ala dipeptide may be present in

aqueous solution 124. Those ROA calculations were performed using restricted Hartree-Fock

theory with a split valence plus basis set, and neglected the effects of explicit water on the

chiroptical response tensor. Pecul et al. showed recently that solvent influences on the chiroptical

Page 52: ABSTRACT - Duke University

52

response tensor should be included when computing solvent effects on ROA 112. Jalkanen et al.

recently computed the ROA spectra of a single (H2O)4-Ala dipeptide cluster, with the PPII

conformation of the dipeptide, embedded in implicit solvent models using the B3PW91/B3LYP

exchange correlation functionals with the aug-cc-pVDZ basis set 125. Those simulated ROA

spectra 124-126 were in qualitative agreement with the experimental spectra, which provided the

impetus for computing the ROA spectrum with improved accuracy for aqueous Ala dipeptide

solutions using geometry sampling as an explicit part of the calculation. We now show that

including explicit solute-solvent interactions is essential for computing the ROA of Ala dipeptide

in water, and that TD-DFT computation (with the RI-J approximation and rDPS basis set) using

MC sampled geometries in aqueous solution provides a suitable means to analyze the ROA

spectra of peptides and determine the dominant populated conformations of peptides in solution.

4.3 Results

4.3.1 Low-energy conformations of Ala dipeptide in water

Figure 4.1 shows the difference in free energy (ΔG) of Ala dipeptide conformations in

water computed using MC simulations. The αR, αL, β, and PPII conformational regions are

populated by the dipeptide (Figure 4.1B), in agreement with previous molecular mechanics force

field simulations 116,121. Figure 4.1C shows that the computed distribution of φ and ψ dihedral

angles (for Ala dipeptide conformations with ΔG < 0) also resembles the φ, ψ distribution of

amino acid residues that are not found in either canonical helix or sheet secondary structures in

protein X-ray crystal structures 127. Ala dipeptide explores both the right- (αR) and left- (αL)

handed helical conformations in agreement with previous MD and MC simulation studies 68,116,121.

Thus, our MC simulations of aqueous Ala dipeptide indicate that the α-helical conformations are

stabilized by solvation.

Page 53: ABSTRACT - Duke University

53

Figure 4.1 . A. Ball and stick model for the N-acetylalanine-N′-methylamide (Aladipeptide). The peptide backbone structure in Ala dipeptide is defined by φ and ψ dihedralangles. B. Free energy differences of Ala dipeptide conformations in water computed using

Monte Carlo simulations. αR, αL, β, and PPII conformational regions are shown on theRamachandran map. Figure 4.1C shows the computed φ and ψ distribution (yellow)

resembles the empirically derived φ, ψ distribution of amino acid residues (black) not ineither canonical helix or sheet secondary structures in protein X-ray crystal structures 127.

Page 54: ABSTRACT - Duke University

54

4.3.2 Structures of H2O-Ala dipeptide clusters

H2O-Ala dipeptide clusters from the MC simulations, corresponding to the low-energy

regions (αR, αL, β, and PPII; Figure 4.1B) of the Ramachandran map, were optimized using

quantum mechanical energy minimization (see computational method section for details). Figure

4.2 shows the Ramachandran map for Ala dipeptide structures taken from the MC simulations

prior to geometry optimization. Each φ, ψ pair (black) represents a cluster of H2O-Ala dipeptide

clusters that have the same peptide backbone dihedral angles with different water geometries. The

φ, ψ values correspond to peptide geometries from low-energy regions 1, 2, and 3 of the

Ramchandran map (Figure 4.2A). The φ, ψ values for H2O-Ala dipeptide structures, and the

number of structures for each φ, ψ pair, appear in Table 1 (360 total structures). Figure 4.2B

shows a wider distribution of φ, ψ values for the structures of H2O-Ala dipeptide clusters,

optimized using quantum mechanical energy minimization. The peptide backbone dihedral angles

changes upon geometry optimization. However, each peptide conformation, before and after

geometry optimization, occupies the same region of the Ramachandran map. Following geometry

optimization of the H2O-Ala dipeptide cluster, the structures were used for computing ROA and

Raman spectra using TD-DFT (see method section for details).

Page 55: ABSTRACT - Duke University

55

Figure 4.2 Ramachandran map of φ, ψ pairs for Ala dipeptide-water clusterstructures taken from the Monte Carlo simulations. Figure 4.2A shows a collection of Ala

dipeptide-water cluster structures (black) that have the same peptide backboneconformation with different water arrangements. Figure 4.2B shows the distribution of φ, ψvalues for the alanine dipeptide-water clusters optimized using quantum mechanical energyminimization. Green arrows show that each peptide conformation, before and after energy

minimization, occupies the same region on the Ramachandran map.

Page 56: ABSTRACT - Duke University

56

Table 4-1 The peptide backbone dihedral angles (φ, ψ) of alanine dipeptide-watercluster structures from the Monte Carlo simulations used in the quantum mechanical

energy minimization.

φ, ψ values

(deg)

Number of structures

with explicit water

-150, 150 30

-140, 135 30

-120, 115 30

-90, 120 30

-120, 150 30

-105, 150 30

-80, 145 30

-60, 165 30

-60, 135 30

-145, 80 30

-65, -45 20

-60, -30 10

55, 65 30

Page 57: ABSTRACT - Duke University

57

4.3.3 ROA and Raman spectra of H2O- and D2O- Ala dipeptideclusters

Figure 4.3 shows the computed ROA (A, C) and the Raman (B, D) spectra using H2O-

Ala dipeptide clusters. The computed spectra are averaged (see discussion for details) over 30

H2O-Ala dipeptide structures from PPII (inset in Figure 4.3A) and αR (inset in Figure 4.3C)

conformational regions of the Ramachandran map. The ROA spectra computed using both the

PPII and αR conformations, taken together, contain all the key spectral features found in the

measured spectrum. For example, ROA bands (a-f in Figure 4.3A) at 395 cm-1 (a), 940 cm-1 (b),

1140 cm-1 (c), 1560 cm-1 (e, amide II), 1665 cm-1 (f, amide I), and the couplet centered at ~1300

cm-1 (d, extended amide III) are predicted in the computed spectrum using a set of Ala dipeptide

conformations with -80° ≤ φ ≤ -60° and 130° ≤ ψ ≤ 150° (blue, inset in Figure 4.3A), which is the

set of structures from the PPII conformational region on the Ramachandran map. Thus, these

ROA bands are attributed to the H2O-Ala dipeptide structures with PPII conformations. The range

of φ, ψ values (-80° ≤ φ ≤ -60° and 130° ≤ ψ ≤ 150°) for the PPII conformations of Ala dipeptide

in aqueous solution determined from our ROA analysis is consistent with the range of values (-

95° ≤ φ ≤ -45° and 95° ≤ ψ ≤ 145°) determined by 2D IR spectroscopy 40. Barron and co-workers

have shown that positive ROA bands at ~1318 cm-1 and ~1670 cm-1 characterize PPII

conformations of polypeptides 42,128, also consistent with our ROA analysis.

Page 58: ABSTRACT - Duke University

58

Figure 4.3 SCP backscattering ROA (A, C) and Raman (B, D) spectra of H2O-Aladipeptide clusters. The computed spectra shown in blue and red are averaged over a

collection of dipeptide conformations from PPII (inset in A) and αR (inset in C) regions ofthe Ramchandran map. The experimental spectra (green; in arbitrary units) are from

reference 126.

Figure 4.3C shows that the ROA bands (a-f) at 395 cm-1 (a), 940 cm-1 (b), 1140 cm-1 (c),

1380 cm-1 (e, Cα-H bend), 1450 cm-1 (f), and the positive-negative couplet centered at ~1300 cm-1

(d, extended amide III) are predicted in the computed spectrum using a set of Ala dipeptide

Page 59: ABSTRACT - Duke University

59

conformations (-50° ≤ φ ≤ -100° and -50° ≤ ψ ≤ 0°) from the αR region of the Ramachandran map

(red, inset in Figure 4.3C). Thus, these ROA bands can be attributed to the H2O-Ala dipeptide

structures with αR conformations of the dipeptide. Interestingly, the ROA spectrum computed

using the set of αR structures show a sequence of -/- bands in the frequency range of 300 cm-1 to

400 cm-1 (Figure 4.3C), consistent with the measured spectra. In contrast, the ROA spectrum

computed using the set of PPII structures show a sequence of +/- bands in the same frequency

range of 300 cm-1 to 400 cm-1 (Figure 4.3A). Thus, the difference in the signs of the predicted

ROA bands for the αR and PPII conformations suggests that the ROA bands in the frequency

range of 300 cm-1 to 400 cm-1 can be used to differentiate between the right-handed αR and left-

handed helical PPII conformations of Ala dipeptide (see discussion section for more detail). The

Raman spectra computed using the PPII (Figure 4.3B) and αR (Figure 4.3D) structures are also in

good agreement with the measured spectrum. Figures 4.3B and 4.3D show only subtle differences

in the computed Raman spectra using the PPII and αR conformations. For example, Raman bands

at ~1140 cm-1 (Figure 4.3B) and ~1540 cm-1 (Figure 4.3D) are predicted for the PPII and αR

conformations, respectively. Thus, ROA is a better probe of the secondary structure of peptides in

aqueous solution than Raman scattering experiments.

Page 60: ABSTRACT - Duke University

60

Figure 4.4 ROA spectra computed using Ala dipeptide-H2O cluster from the low-energy region of the Ramachandran map. The computed spectra shown are averaged over a

collection of dipeptide conformations from regions A, B, and C of the Ramchandran map(inset in A). The experimental spectrum (in arbitrary units) is from reference 126.

Figure 4.4 and 4.5 show a more detailed analysis of the dependence of the computed

ROA spectra on the φ and ψ dihedral angles of Ala dipeptide in H2O. The computed spectra

correspond to Ala dipeptide-water clusters from the low-energy regions of the Ramchandran map.

For example, Figures 4.4 A, 4.5 E and 4.5 F show the computed ROA spectra using sets of Ala

dipeptide conformations from the β, PPII and αR conformational region of the Ramchandran map

(insets in Figures 4.4 and 4.5), respectively. The computed spectra are averaged over sets of Ala

Page 61: ABSTRACT - Duke University

61

dipeptide conformations on 50° × 50° grids of φ and ψ. The ROA spectral features in the

computed spectra using peptide-water clusters with the PPII (Figure 4.5E), αR (Figure 4.5F) and

β (Figure 4.4A) conformations of Ala dipeptide are very different, which originate from different

peptide and water geometries of the Ala dipeptide-water clusters.

Page 62: ABSTRACT - Duke University

62

Figure 4.5 ROA spectra computed using Ala dipeptide-H2O cluster from the low-energy region of the Ramachandran map. The computed spectra shown are averaged over acollection of dipeptide conformations from regions D, E, F and G of the Ramchandran map

(inset in D). The experimental spectrum (in arbitrary units) is from reference 126.

Page 63: ABSTRACT - Duke University

63

Figure 4.6 SCP backscattering ROA (A, C) and Raman (B, D) spectra of D2O-Aladipeptide clusters. The computed spectra shown in blue and red are averaged over a

collection of dipeptide conformations from PPII (inset in A) and αR (inset in C) regions ofthe Ramchandran map. The experimental spectra (green; in arbitrary units) are from

reference 126.

Figure 4.6 shows the computed ROA (A, C) and the Raman (B, D) spectra using D2O-

Ala dipeptide clusters. The computed ROA spectra using both the PPII and αR conformations

(insets in Figure 4.6A and 4.6C) taken together contain all the key spectral features found in the

Page 64: ABSTRACT - Duke University

64

measured spectrum in D2O. For example, ROA bands (peaks a-h in Figure 4.6A) at 340 cm-1 (a),

395 cm-1 (b), 500 cm-1 (c), 950 cm-1 (d), 1050 cm-1 (e), 1280 cm-1 (f, extended amide III), 1440

cm-1 (g), and 1650 cm-1 (h, amide I), are predicted in the computed spectrum using the set of PPII

conformations. Similarly, ROA bands (peaks a-e in Figure 4.6C) at 340 cm-1 (a), 395 cm-1 (b),

500 cm-1 (c), 1150 cm-1 (d), and 1440 cm-1 (e) are predicted in the computed spectrum using the

set of αR conformations. Interestingly, the measured ROA spectrum of Ala dipeptide in D2O

shows a sequence of +/-/- bands in the frequency range of 300 cm-1 to 400 cm-1. In contrast, the

measured ROA spectrum in H2O (Figure 4.3) shows a sequence of -/- bands between 300 cm-1

and 400 cm-1. Thus, the ROA spectrum of Ala dipeptide in D2O has an additional positive band in

the frequency range of 300 cm-1 to 400 cm-1. The ROA spectra computed using the set of PPII

(Figure 4.6A) and αR (Figure 4.6C) structures in D2O show a sequence of +/(-) and -/(-) bands in

the frequency range of 300 cm-1 to 400 cm-1, respectively. The sign in the parenthesis, (-),

indicates the common negative ROA band predicted in the frequency range of 300 cm-1 to 400

cm-1 for the PPII and αR conformations. Thus, the computed ROA spectra based on PPII and αR

dipeptide conformations taken together contain the +/(-)/- spectral feature found in the frequency

range of 300 cm-1 to 400 cm-1 of the measured spectrum in D2O. The Raman spectra computed

using the PPII (Figure 4.6B) and αR (Figure 4.6D) structures are also in good agreement with the

measured spectrum in D2O.

Page 65: ABSTRACT - Duke University

65

Figure 4.7 ROA spectra computed using Ala dipeptide-D2O cluster from the low-energy region of the Ramachandran map. The computed spectra shown are averaged over a

collection of dipeptide conformations from regions A, B and C of the Ramchandran map(inset in A). The experimental spectrum (in arbitrary units) is from reference 126.

Page 66: ABSTRACT - Duke University

66

Figure 4.8 ROA spectra computed using Ala dipeptide-D2O cluster from the low-energy region of the Ramachandran map. The computed spectra shown are averaged over acollection of dipeptide conformations from regions D, E, F and G of the Ramchandran map

(inset in D). The experimental spectrum (in arbitrary units) is from reference 126.

Page 67: ABSTRACT - Duke University

67

Figure 4.7 and 4.8 show a more detailed analysis of the dependence of the computed

ROA spectra on the φ and ψ dihedral angles of Ala dipeptide in D2O. The computed spectra

correspond to Ala dipeptide-water clusters from the low-energy regions of the Ramchandran map.

The ROA spectral features in the computed spectra using peptide-water clusters with the PPII

(Figure 4.8E), αR (Figure 4.8F) and β (Figure 4.7A) conformations of Ala dipeptide in aqueous

solution are very different. Thus, similar to the H2O results, different ROA spectra originate from

very different peptide and water geometries of the peptide-water clusters.

Page 68: ABSTRACT - Duke University

68

Figure 4.9 SCP backscattering ROA (A, C) and Raman (B, D) spectra of Aladipeptide in H2O. The computed spectra shown in orange and pink are averaged over a

collection of dipeptide conformations from αL (inset in A) and β (inset in C) regions of theRamchandran map. The experimental spectrum (in arbitrary units) is from reference 126.

Page 69: ABSTRACT - Duke University

69

Figure 4.10 SCP backscattering ROA (A, C) and Raman (B, D) spectra of Aladipeptide in D2O. The computed spectra shown in orange and pink are averaged over a

collection of dipeptide conformations from αL (inset in A) and β (inset in C) regions of theRamchandran map. The experimental spectrum (in arbitrary units) is from reference 126.

Figures 4.9 and 4.10 show that there is no significant similarity between the measured

and computed ROA and Raman spectra using the αL (with 45° ≤ φ ≤ 65° and 25° ≤ ψ ≤ 55°) and

β (with -180° ≤ φ ≤ -125° and 150° ≤ ψ ≤ 180°) conformations of H2O-Ala dipeptide and D2O-

Page 70: ABSTRACT - Duke University

70

Ala dipeptide clusters. Thus, our ROA analysis suggests that Ala dipeptide in water populates αR

and PPII conformations without substantial αL and β populations.

Figure 4.11 ROA (A, C) and Raman (B, D) scattering cross sections contributions ofAla dipeptide (red), water (blue), and peptide-water interactions (black) to the total ROA

and Raman scattering cross sections of Ala dipeptide-water clusters (green). Thesecomputed spectra are averaged over a collection of dipeptide conformations from PPII (A,

B) and αR (C, D) regions of the Ramachandran map.

Figure 4.11 shows the relative contributions of the peptide, water and peptide-water

coupling to the ROA (A, C) and Raman (B, D) intensities computed using the αR (A, B) and PPII

Page 71: ABSTRACT - Duke University

71

(C, D) conformations of Ala dipeptide. The ROA and Raman scattering cross sections are

influenced by water in the low frequency range of 300 cm-1-1100 cm-1 and for the amide I and II

bands. However, water makes no contribution to the Raman and ROA intensities from 1200 cm-1

to 1500 cm-1. Kapitán et al. recently showed that ROA bands in the 800 cm-1-1600 cm-1 range for

the proline zwitterion in aqueous solution are not influenced by water 129, which is consistent with

our ROA analysis that also predicts the presence of ROA bands (1200 cm-1-1500 cm-1) for Ala

dipeptide in aqueous solution that are not influenced by the solvent.

Figure 4.12 SCP backscattering ROA spectra computed for a representative PPII (φ= -68° and ψ = 135°; blue) and a αR conformation of Ala dipeptide (φ = -73° and

ψ = −30°; red). Experimental spectra (green; in arbitrary units) in H2O (A, B) and D2O (C,D) are from reference 126.

Page 72: ABSTRACT - Duke University

72

Figure 4.12 shows the dipeptide contributions of the total ROA and Raman scattering

cross sections of Ala dipeptide-water cluster (see equations 5 and 6 in the method section for

more detail), computed using a PPII (φ = -68 and ψ = 135°; A, C) and a αR (φ = -73° and ψ = -

30°; B, D) conformation of Ala dipeptide. The computed ROA spectrum using both of these

conformations show spectral features found in the measured spectrum in both H2O (A, B) and

D2O (C, D). For example, Figures 4.12B and 4.12D show that ROA bands a-i (including amide I,

II III and the low frequency vibrational modes in the 300 cm-1 to 400 cm-1 range) predicted in the

computed spectra using the αR conformations of Ala dipeptide conformation are observed in the

measured spectra in H2O and D2O.

In order to understand the origin of the ROA intensity differences associated with the

molecular vibrations in the low frequency range (300 cm-1-400 cm-1), we decomposed those

intensity differences into contributions from groups of atoms in Ala dipeptide. As proposed by

Hug 19,130, the ROA intensity decompositions is illustrated using the group coupling matrices for a

PPII (φ = -68 and ψ = 135°; Figure 4.13A) and an αR (φ = -73° and ψ = -30°; Figure 4.13B)

conformation of Ala dipeptide (Figure 4.13I). The group coupling matrices show that the ROA

intensity differences at ~364 cm-1 (Figure 4.13C) for the PPII (Figure 4.13A) and at ~325 cm-1

(Figure 4.13D) for the αR (Figure 4.13B) are both dominated by the vibrational coupling of

groups 1 and 2 (Figure 4.13I), which is positive (red) and negative (yellow) for the PPII and αR

conformations, respectively. Figure 4.13 also show that the ROA intensity differences associated

with molecular vibrations at ~401 cm-1 (13G) for the PPII and at ~395 cm-1 (13H) for the αR are

both dominated by the vibrational couplings of groups 1 and 3 (Figure 4.13I), and is negative for

both of these conformations. Thus, the ROA spectral decomposition describe the physical basis

for the pattern of +/− and −/− ROA bands in the frequency range of 300 cm-1 to 400 cm-1 that are

predicted for the PPII and αR helical conformations of Ala dipeptide, respectively. Figures 4.13E

and 4.13F show the ROA intensity differences associated with molecular vibrations at ~352 cm-1

for the PPII and at ~338 cm-1 for the αR conformation.

Page 73: ABSTRACT - Duke University

73

Figure 4.13 ROA intensity differences associated with the vibrations in the lowwavenumber range decomposed into contributions from groups of atoms in Ala dipeptidefor a PPII (φ = -68° and ψ = 135°; A) and a αR (φ = -73° and ψ = −30°; B) conformation of

Ala dipeptide. The groups of atoms in Ala dipeptide are shown in panel I. Positive andnegative ROA intensity differences are shown as red and yellow circles, respectively.

Page 74: ABSTRACT - Duke University

74

4.4 Discussion

The aim of this study is to characterize Ala dipeptide conformations that are consistent

with the measured ROA spectra in H2O and D2O. The ROA spectra were computed for H2O- and

D2O-Ala dipeptide clusters with the PPII, αR, αL and β conformations of the dipeptide (Figure

4.1B) because the MC simulations found these structures in aqueous solution. Agreement

between the measured and averaged computed ROA spectra (based on overlay of spectra) is used

to characterize the populated conformations of Ala dipeptide in aqueous solution. The computed

ROA spectra for the H2O- and D2O-Ala dipeptide clusters with the PPII and αR conformations

contain spectral features that are observed in the measured spectra (Figures 4.3 and 4.6).

However, there is no significant similarity between the measured and predicted ROA spectra for

the αL and β conformations. Thus, our analysis of ROA of Ala dipeptide in aqueous solution

suggests that Ala dipeptide populates both the αR and PPII conformations, but no substantial

population of αL and β structures, despite the fact that αL and β structures are found in the MC

simulations.

The computed “average” ROA spectra are arithmetic averages, which implies that we are

assuming an equal population of Ala dipeptide structures within the PPII, αR, αL and β

conformational regions on the Ramachandran map (insets in Figures 4.3, 4.6 and 4.9). In other

words, the φ, ψ pairs of Ala dipeptide structures within the PPII, αR, αL and β conformational

regions are assumed to be equally probable. A quantitative comparison of the predicted and

observed ROA spectra requires the Boltzmann distribution of Ala dipeptide conformations, and

the population-weighted average computed ROA spectrum. In principle, we could use the energy

of each Ala dipeptide-water cluster (from the DFT geometry optimization calculation, see

methods for details) to determine the Boltzmann distribution of Ala dipeptide-water cluster

geometries, and hence a population-weighted average computed ROA spectrum. The

experimental observable, i.e. the ROA spectrum of Ala dipeptide in aqueous solution, is an

ensemble average over both solvent positions and peptide geometries. However, each Ala

dipeptide-water cluster used in our ROA calculations constitutes specific peptide-water geometry,

Page 75: ABSTRACT - Duke University

75

and not an ensemble average. Hence, a population-weighted average computed using the set of

Ala dipeptide-water clusters would still not be an accurate prediction of the experimental

observable. Thus, we resort to the assumption of an equal probability of Ala dipeptide structures

within the PPII, αR, αL and β conformational regions, and compare the computed average ROA

spectrum of each of these conformational regions to the observed spectrum.

The conformational heterogeneity of Ala dipeptide has been probed using other

spectroscopic methods. For example, the 13C NMR spectra of Ala dipeptide in water and in

liquid-crystalline media suggest that its structures in aqueous solution and in the liquid-crystalline

state are very similar, and that the structure is dominated by PPII conformations 37,38. However,

these studies could not rule out the existence of αR helical conformations in water. Following

these NMR studies, Mehta et al. used 13C NMR to show that Ala dipeptide in water exists in a

mixture of PPII and αR conformations 39, consistent with our ROA analysis. Kim et al. used 2D-

IR to probe the conformational heterogeneity of Ala dipeptide in aqueous solution 40. Those

authors concluded that the 2D-IR experimental data and simulations are consistent with PPII-like

conformations of Ala dipeptide in aqueous solution, and ruled out the possibility of αR helical

conformations for Ala dipeptide in water. We emphasize that the 2D-IR experiments were

performed in D2O, which stabilizes PPII conformations compared to H2O. For example,

Chellgren and Creamer have also shown that D2O stabilizes PPII conformations of alanine

containing peptides relative to H2O 131. Those authors therefore suggested that conformational

analysis of peptides using NMR, VCD, IR, and Raman spectroscopy of peptides in D2O is biased

towards increased PPII populations.

Deng et al. suggested that the two negative ROA bands observed in the frequency range

of 300 cm-1 to 400 cm-1 in the measured spectra of Ala dipeptide in H2O and D2O are indicative of

its predominant conformation in aqueous solution 126. Our ROA computations predict the

sequence of +/- and -/- ROA bands in the frequency range of 300 cm-1 to 400 cm-1 for the PPII

and αR helical conformations of Ala dipeptide, respectively (Figure 4.12). Thus, the ROA

analysis suggests that the two negative ROA bands observed in the frequency range of 300 cm-1

Page 76: ABSTRACT - Duke University

76

to 400 cm-1 in the measured spectra in H2O and D2O are characteristics of αR helical

conformations of Ala dipeptide. The measured ROA spectrum of Ala dipeptide in D2O (Figures

4.12C and 4.12D) also shows a positive ROA band at ~320 cm-1 in addition to the two negative

peaks in the frequency range of 300 cm-1 to 400 cm-1. Our ROA spectrum computed using the

PPII conformation of Ala dipeptide in D2O (Figures 4.12C) predicts a positive ROA band in the

frequency range of 300 cm-1 to 400 cm-1, which consistent with the measured ROA spectrum in

D2O. Thus, the positive ROA band in the frequency range of 300 cm-1 to 400 cm-1 is a

characteristic of the PPII conformations of Ala dipeptide. Since, D2O stabilizes PPII

conformations of peptides relative to H2O (see discussion in preceding paragraph), the positive

ROA band observed in D2O, is not observed in the measured spectrum in H2O. The predicted

ROA spectra using PPII and αR helical conformations of Ala dipeptide have common spectral

features found in the measured spectra in H2O and D2O, which indicate that Ala dipeptide in

aqueous solution displays both conformations. We emphasize that the differential ROA response

of Ala dipeptide in H2O and D2O, such as the key experimental spectral feature of -/- and +/-/-

bands in the frequency range of 300 cm-1 to 400 cm-1 in H2O and D2O, respectively, predicted by

our calculations demonstrates that ROA can also differentiate the PPII and αR dominant

populated structures of Ala dipeptide in aqueous solution.

Han et al. also computed the ROA spectra of Ala dipeptide in H2O and their theoretical

analysis of the ROA spectra suggests that PPII is the dominant conformation in aqueous solution124. In contrast, our theoretical analysis of the ROA spectra of Ala dipeptide suggests that αR is the

dominant conformation in H2O. If PPII dominated the conformational ensemble of Ala dipeptide

in H2O then a positive ROA band in the frequency range of 300 cm1 to 400 cm-1 would

characterize it (see discussion in preceding paragraph). However, the measured ROA spectrum in

H2O (Figure 4.12B) shows two negative bands in that frequency range. The simulated ROA

spectrum based on the H2O-Ala dipeptide cluster with the αR conformation of the dipeptide

(Figure 4.12B) shows negative ROA bands between 300 cm-1 and 400 cm-1, in agreement with the

measured spectrum. Thus, our ROA calculations suggest a dominant population of αR

Page 77: ABSTRACT - Duke University

77

conformation of Ala dipeptide in H2O, which is also consistent with 13C NMR studies 39. The

measured ROA spectrum of Ala dipeptide in D2O (Figure 4.12C) shows a positive ROA band

between 300 cm-1 and 400 cm-1 that is characteristic of the PPII conformation (see discussion in

preceding paragraph), which is accurately predicted in our ROA computation using D2O-Ala

dipeptide cluster with the PPII conformation of the dipeptide (Figure 4.12C).

Similarities between the observed and computed spectra (Figures 4.3, 4.6, 4.12) using the

PPII and αR conformations of Ala dipeptide suggest that the pattern of +/− and −/− ROA bands in

the frequency range of 300 cm-1 to 400 cm-1 are characteristic of the PPII and αR helical

conformations for Ala dipeptide, respectively. In order to assign these spectral features to

molecular structural elements, we decomposed the ROA intensity differences associated with

vibrations in the frequency range of 300 cm-1 to 400 cm-1 into contributions from nuclei in groups

of atoms in Ala dipeptide for a αR and a PPII conformation. For example, the ROA band at ~364

cm-1 (Figure 4.13A) and ~325 cm-1 (Figure 4.13B) for the PPII and αR, respectively, both

originate from the relative motions of the nuclei in groups 1 and 2 (Figure 4.13I). ROA bands that

are dominated by the relative motions of the nuclei in groups 1 and 2 in Ala dipeptide (Figure

4.13I) are also a probe of its secondary structure, since both of these groups contain the planar

amide bonds that determine the φ and ψ dihedral angles. Thus, the ROA spectral features in the

low frequency range of 300 cm-1 to 400 cm-1 provide structural information of the predominant

conformations of Ala dipeptide in aqueous solution.

The ROA solvent dependence for chiral molecules was computed recently using DFT.

Pecul et al. used DFT combined with a continuum solvent description to examine the influence of

solvent on the ROA spectra of rigid molecules 112. The authors concluded that the solvent

influence on the molecular geometry, the molecular Hessian, and the optical tensors should be

included when computing solvent effects on ROA. Our computation of the geometry

optimization, vibrational frequencies, normal modes, and derivatives of electric dipole-electric

dipole polarizability and electric dipole-magnetic dipole polarizability tensors (Equation 4 in

method section) were carried out using Ala dipeptide-water cluster structures. Thus, including

Page 78: ABSTRACT - Duke University

78

explicit water in our ROA calculations accounts for the necessary solvent effects. We also find

that the ROA and Raman scattering cross sections in the low frequency range of 300 cm-1-1100

cm-1 and for the amide I and II bands are influenced by water, in contrast, the ROA and Raman

bands in the frequency range of 1200 cm-1 to 1500 cm-1 are not affected by water (Figure 4.11).

This result is consistent with a recent ROA study of proline zwitterion in aqueous solution, which

also showed the presence of ROA bands that are not influenced by water 129. ROA computation of

polypeptides including explicit solute-solvent interactions becomes extremely time consuming as

the number of atoms grows. Efficient ROA computations of polypeptides may be achieved by

excluding the water contribution to the Cartesian gradients of the electric dipole-electric dipole

polarizability and electric dipole- magnetic dipole polarizability. However, the solvent influence

on the geometry optimization of the solute and on the molecular Hessian must be included for

predicting of the ROA in solution.

The accurate prediction of ROA spectra of peptides in aqueous solution requires

averaging over both solvent positions and peptide geometries 108. The average ROA spectra

(Figures 4.3, 4.6) computed using H2O- and D2O-Ala dipeptide clusters with the PPII and αR

peptide conformations predict ROA spectral features of Ala dipeptide in aqueous solution.

Figures 4.12B and 4.12D show that the ROA spectra computed using a single peptide

conformation with φ = -73° and ψ = -30° agrees better with the observed spectrum than the

computed spectrum averaged over a collection of Ala dipeptide structures with a wide range of φ

and ψ values (-50° ≤ φ ≤ -100° and -50° ≤ ψ ≤ 0°; Figures 4.3C, 4.6C) from the αR

conformational region of the Ramachandran map. This suggests that a better agreement between

the observed and computed average ROA spectrum of the αR conformational region might be

achieved using a narrow distribution of peptide geometries around φ = -73° and ψ = -30°, rather

than the broader distribution of structures with -50° ≤ φ ≤ -100° and -50° ≤ ψ ≤ 0°.

Our calculations show that the predicted ROA spectra using both the PPII and αR

conformations of Ala dipeptide in H2O and D2O are in very good agreement with the measured

spectra (Figures 4.3 and 4.6) based on the overlap between the measured and computed spectra.

Page 79: ABSTRACT - Duke University

79

Herrmann et al. recently showed that the ROA of helical peptides are dominated by the chirality

associated with the peptide backbone conformation 113,132. Thus, based on the good agreement

between the measured and computed spectra using the helical PPII and αR conformations of Ala

dipeptide, we suggest that the backbone conformation of Ala dipeptide in aqueous solution is

characterized by the compact αR and extended PPII conformations, both of which are present in

aqueous solution.

Conformational analysis of Ala dipeptide is of fundamental importance to the question of

the local conformational preference of unfolded polypeptides. Our ROA computational method

could be used to interpret the ROA spectra of longer unfolded polypeptides in aqueous solution,

and to characterize their solution conformations. In order to characterize conformations of

unfolded polypeptides, one has to determine residue-specific Ramachandran maps of the

polypeptides. We demonstrate that ROA probes of molecular chirality can be used to determine

the Ramachandran map of the shortest alanine-containing peptide. The accurate description of the

conformational heterogeneity of Ala dipeptide backbone is an important step towards defining the

residue-specific conformational preferences in unfolded polypeptides. In addition to the

biological significance of dipeptide conformational analysis, combined experiments and

simulations on short peptides can be used to determine the accuracy of commonly used force

fields in MD and MC simulations. For example, our MC and ROA studies of Ala dipeptide show

that the shortest alanine-containing peptide exists as a mixture of PPII and αR conformations in

aqueous solution, without substantial population of left-handed αL or β structures, despite

sampling αL and β structures in force field based MC simulations. Best et al. recently studied

polyalanine peptide A5 conformations in aqueous solution using MD simulations in order to

address the question of whether the commonly used force fields overpopulate the α-helical

conformations of polyalanine peptides102. They concluded that the NMR data are consistent with

force fields that produce α-helical populations of polyalanine peptides and that the data do not

require exclusive formation of the PPII structure, consistent with our conformational analysis of

Ala dipeptide in aqueous solution.

Page 80: ABSTRACT - Duke University

80

Growing access to commercial ROA spectrometers (http://www.btools.com) and

improved efficiency of ROA calculations 133, such as implementation of analytical derivative

procedures for the computation of ROA, are likely to motivate further ROA analysis of aqueous

polypeptide solutions. A combined conformational analysis and ROA study of polypeptides

should be helpful for understanding more fully the relationship between peptide conformations

and ROA spectra. A link of this kind may provide further insights into the connection between

structure and dynamics.

4.5 Conclusions

Backscattered circular polarization ROA analysis for Ala dipeptide based on TD-

DFT/DFT calculations and MC sampled peptide-water clusters were used with experimental

ROA and Raman data to assign the dominant conformations of Ala dipeptide in aqueous solution.

Agreement of observed and simulated ROA spectral features indicate the viability of screening

plausible MC structures based on ROA data. We find that the computed ROA spectra using a

PPII conformations with φ = -68° and ψ = 135° and an αR conformations with φ = -73° and ψ = -

30° are in agreement with the observed spectra. Thus, our ROA analysis shows that Ala dipeptide

adopts both αR and PPII conformations in aqueous solution, which are the dominant

conformations in aqueous solution.

We also find that including explicit solute-solvent interactions is required for: (a)

adequate quantum mechanical geometry optimization of the solute conformations in solution, (b)

computation of the molecular Hessian and (c) computation of the solvent influence on the optical

tensor for predicting the ROA spectra of peptides in solution. The qualitative agreement between

the observed and computed spectra using the PPII and αR conformations of the dipeptide in H2O

and D2O suggest that our ROA computational method could be used to interpret measured ROA

spectra of longer polypeptides in aqueous solution and to characterize their solution

conformations.

Page 81: ABSTRACT - Duke University

81

4.6 Methods

Two sets of MC computer simulations of Ala dipeptide in H2O were performed using the

OPLS force field 83 and TIP4P water model 134. First, low-energy conformations of Ala dipeptide

in H2O were identified that characterize potentially populated regions of the Ramachandran map

using MC free-energy simulations. Second, H2O configurations were sampled with fixed

backbone dihedral angles (φ and ψ; Figure 4.1A) for the dipeptide. The fixed peptide geometries

were selected from low-energy regions of the Ramachandran map. H2O-Ala dipeptide cluster

structures from the second set of MC simulations (sampling water configuration with fixed-

peptide geometry) were optimized using DFT. Each H2O-Ala dipeptide cluster has ten water

molecules within a cut-off distance of 0.5 nm from the hydrogen bond donor (NH) and acceptor

(O, N) atoms of Ala dipeptide (Figure 4.1A). Following geometry optimization of the H2O-Ala

dipeptide cluster, the structures were used for computing the back-scattering ROA and Raman

spectra using TD-DFT with RI-J approximation, the BLYP functional and the rDPS basis set.

4.6.1 MC simulations

Free energy perturbation theory 88 was used to calculate the potential of mean force

(PMF) that describes the relative free energy difference between Ala dipeptide conformations as a

function of the peptide backbone dihedral angles. One-dimensional (1D) PMFs were calculated

first with φ fixed and ψ varied from 0° to 360° in 5° increments. Then ψ was fixed and φ was

varied from 0° to 360° in 5° increments. A 2D plot of the free energy difference for φ = 80°, ψ =

80°, the reference conformation, and all φ, ψ pairs were calculated from the 1D PMFs. Thus, the

2D plot identifies low-energy conformations of Ala dipeptide in water.

MC simulations of Ala dipeptide in a box of 500 water molecules were performed using

BOSS 88 at 25°C and 1 atm (NPT ensemble) with periodic boundary conditions. Long-range

electrostatic interactions were evaluated by the Ewald method 135. The AM1 91 based CM1A 92

partial atomic charges for the solute and solvent were used in the MC simulations. A more

detailed description of the MC move and the atomic charge-partitioning scheme appears in 88.

Page 82: ABSTRACT - Duke University

82

4.6.2 TD-DFT calculations of ROA

Zuber et al. recently demonstrated accurate and efficient prediction of ROA spectra

using DFT 17. The approach avoids calculating the four-center Coulomb integrals and omits

contributions of the electric dipole-electric quadrupole polarizability tensor to the ROA

differential scattering cross-section. Luber et al. also showed recently that the contribution of the

electric dipole-electric quadrupole polarizability tensor to the ROA intensities can be neglected

for the amide I, amide II, amide III vibrational modes for peptides (below 2000 cm-1) 136. We use

the same scheme in the off-resonance approximation 16 to calculate the circular difference

differential scattering cross-sections per unit solid angle

−Δndσ π( )SCPdΩ

for back scattered

circular polarization 137 (SCP) with naturally polarized (n) incident light:

−Δndσ π( )SCPdΩ

≈Kc48βG

2( ) (1)

c is the speed of light,

βG2 is the anisotropic ROA invariant , and K is:

K =107

9π 2µo

2c 4 ˜ ν oωo

200πc−Δ ˜ ν p

3

(2)

ωo = 2πν o is the frequency of the incident light,

Δ ˜ ν p is the Raman frequency shift (in cm-1), c is

the speed of light (in m s-1), and

µo is the permittivity of vacuum.

Placzek polarizability theory writes 19

βG2 =

12

3αµν ′ G µν −αµµ ′ G νν( )µν

∑ (3)

where

αµν ′ G µν is

αµν ′ G µν =h

8π 2cΔ ˜ ν p i, j∑

∂αµνe

∂xiα

0α ,β∑

∂ ′ G µνe

∂x jβ

0

Lαi,px Lβj ,p

x (4)

Page 83: ABSTRACT - Duke University

83

Lαi,px is the ith component of the Cartesian displacement vector of nucleus α for normal mode p.

αµνe and

′ G µνe are the elements of the electric dipole-electric dipole polarizability tensor and the

imaginary part of the electric dipole-magnetic dipole polarizability tensor, respectively. The index

0 indicates that the derivatives of the electronic tensors are computed at the equilibrium nuclear

geometry.

The geometries of H2O-Ala dipeptide clusters from the second set of MC simulations

were optimized using DFT implemented in Turbomole (Version 5.6) 85 with the BLYP

correlation-exchange functional and the SVP basis set. The normal coordinates of the H2O-Ala

dipeptide clusters were computed from the molecular Hessian, which was determined for the

optimized geometry. Optimizations of some of the H2O-Ala dipeptide cluster structures lead to

imaginary frequencies and hence such structures were not included in the ROA calculations. 532

nm incident laser frequency was used for tensor calculations, which were performed using TD-

DFT with the BLYP functional, the rDPS basis set, and the [7s3p3d1f]/[3s2p1d] auxiliary basis

set 138. All of the calculations were performed in a gauge-independent dipole-velocity formulation

using the RI-J approximation. The Cartesian gradients of the polarizability tensors were evaluated

numerically using the 6N-displaced geometries, where N is the number of atoms. The 6N

geometries were obtained from the optimized geometries of the H2O-Ala dipeptide clusters using

a displacement of ± 10-3 atomic units along the Cartesian coordinates.

The molecular hessian of the optimized geometry of H2O-Ala dipeptide clusters were

used to compute the vibrational frequencies and normal coordinates of D2O-Ala dipeptide

structures. Thus, we used the H2O-Ala dipeptide structures from the MC simulations to compute

the ROA and Raman spectra of D2O-Ala dipeptide clusters. The ROA and Raman spectra of the

D2O-Ala dipeptide clusters were then computed using equation 4.

The total Raman

(ndσ ) and ROA

(−Δndσ ) scattering cross sections of a Ala dipeptide-

water cluster is the sum of scattering cross sections of the peptide

(ndσ p ,−Δndσ p ) , water

(ndσW ,−ΔndσW ) , and peptide-water interactions

(ndσ pWI ,−Δndσ pWI ) . Hence, the total Raman

and ROA scattering cross sections is:

Page 84: ABSTRACT - Duke University

84

ndσ=ndσ p+ndσW +ndσ pWI (5)

−Δ ndσ = −(Δ ndσ p +Δ ndσW +Δ ndσ pWI ) (6)

Thus, the relative contributions of the Raman and ROA scattering cross sections of the

dipeptide, water, and peptide-water interactions can be compared to the total ROA and Raman

scattering cross sections of Ala dipeptide-water cluster.

Page 85: ABSTRACT - Duke University

85

5 Epilogue

In this dissertation we describe computer simulations of the solvent effects on optical

rotation (OR) and Raman optical activity (ROA) of molecules. Our aim is to compute molecular

chiroptical signatures so that the data may be used in combination with experiments to assign and

interpret the conformations and configurations of molecules.

Theoretical OR predictions depend strongly on the quantum mechanical (QM) methods

used, and the results are often strongly geometry dependent. The significant variation in predicted

OR values (sign and magnitude) arise from the small Tr [G′(ω)] (see equations 1 and 2), which is

1-3 order of magnitude smaller than the largest tensor element G′αβ. Thus, in a QM calculation,

changing basis set, level of electron correlation, solvent, molecular geometry, or treatment of

vibrational contributions that slightly change G′ tensor elements can lead to sign inversion in OR.

Hence, theoretical predictions of OR demands QM calculations with different combinations of

basis sets and level of electron correlation in order to check the consistency of the predictions.

Zuber, Beratan and Wipf recently showed that the anisotropic component of Rayleigh optical

activity (RayOA) scattering has a weak dependence on the theoretical methods 139. Hence,

accurate RayOA calculations combined with experiments could be developed into a powerful

chiroptical spectroscopic tool to probe molecular stereochemistry that would complement existing

methods.

Our theoretical studies of methyloxirane in benzene show that the solute’s chiral solvent

imprint can dominate the OR in solution. It will be interesting to determine if chirality transfer

from a chiral solute to an achiral solvent in isotropic phase is a general effect that also occurs in

other chiral systems or for other chiroptical properties. For example, Xu and coworkers recently

showed chirality transfer from methyl lactate 140,141 and methyloxirane 48 to hydrogen-bonded

water using vibrational circular dichroism spectroscopy. Chiral solvent structure around a chiral

solute can influence chemical processes such as asymmetric organic synthesis and chiral

chromatography. Chiral chromatographic techniques are the principal methods for the analysis

and separation of chiral organic and biomolecules, as evidenced by their extensive use throughout

Page 86: ABSTRACT - Duke University

86

the scientific community. However, despite such wide spread use, the molecular mechanisms of

chromatographic separation processes are poorly understood. For example, the chiral solvent

imprint of the chiral surface (stationary phase) could influence the stereoselectivity of the

separation process, and hence, the retention characteristic of solute molecules. A detailed

understanding of the molecular mechanisms of chiral chromatographic separation processes

would help in predicting retention properties of any solute, which can also provide insights into

developing combinations of stationary-mobile phase systems with improved retention

characteristics.

The solution conformations of polypeptides are the subject of increasing attention,

although their ensemble of geometries is difficult to characterize 94. Recent studies suggest that

ROA can be used to characterize backbone conformations of polypeptides in aqueous solution.

ROA is especially useful in cases when assignment of polypeptide conformations by electronic

circular dichroism is difficult, as has been reported recently 142. We show that ROA calculations

based on density functional theory and Monte Carlo (MC) simulations can be used to identify

populated conformations of a peptide in aqueous solution. These calculations facilitate

assignment of the observed ROA spectral features to corresponding secondary structures that

characterize the populated conformations of the peptide in aqueous solution. Thus, the

interpretation of the experimental data enabled by the calculations helps in understanding the

relationship between peptide conformations and ROA spectra. In addition to the agreement

between the theoretical results and the experimental data, our calculations also provide substantial

predictive value. For example, our calculations show that ROA can be used to distinguish

between left-(αL) and right-handed (αR) helical conformations of a peptide backbone in aqueous

solution. To the best of our knowledge this is the first evidence in support of ROA to distinguish

the chirality associated with the handedness of helical conformations of peptides.

The accurate and efficient prediction of ROA suggests that ROA analysis could be used

to identify ROA spectral features that characterize secondary structures (αR, αL, β, PPII) in

proteins and polypeptides. Arora and coworkers recently reported the atomic structure of a short

αR polypeptide stabilized by a main chain hydrogen bond surrogate 143, and Wennermers and

Page 87: ABSTRACT - Duke University

87

coworker showed that Azidopolyproline polypeptide have a single dominant PPII conformation

in aqueous solution 144. ROA analysis of those polypeptide sequences can be used to identify

ROA fingerprints of the αR, β, and PPII conformations. Characterization of the peptide backbone

conformations of intrinsically disordered proteins (IDPs), which do not fold into well-defined

three-dimensional structure under physiological conditions, is the subject of intense interest as

their disorder structures are implicated in important biological functions in cell regulation and

signaling 145,146. The intrinsic disorder of IDPs restricts the utility of X-ray and NMR techniques

for probing their solution conformations. Thus, the ensemble of conformations of IDPs is difficult

to characterize. MC or MD simulations combined with ROA analysis of IDPs may provide the

means to probe for secondary structures of IDPs, and thus, characterize their solution

conformations.

Page 88: ABSTRACT - Duke University

88

Bibliography

(1) Kauffman, G. B.; Myers, R. D. Pasteur's resolution of racemic acid: Asesquicentennial retrospect and a new translation. Chem. Educ. 1998, 3, 1-18.

(2) Oloane, J. K. Optical-activity in small molecules, non-enantiomorphous crystals,and nematic liquid-crystals. Chem. Rev. 1980, 80, 41-61.

(3) Lebel, J. A. Relations which exist between atomic formulas of organic-bodiesand the rotational power of their dissolutions (Reprinted). Bull. Soc. Chim. Fr. 1995, 132,6-16.

(4) Weyer, J. 100 years of stereochemistry - principal development phases inretrospect. Angew. Chem. Int. Ed. 1974, 13, 591-598.

(5) Caldwell, D. J.; H., E. In The theory of optical activity; Wiley-Interscience: NewYork, 1971.

(6) Charney, E. In The molecular basis of optical activity: optical rotatorydispersion and circular dichroism; John Wiley & Sons Inc.: Cambridge, 1979.

(7) Eyring, H.; Kimball, G. E.; Walter, J. E. In Quantum Chemistry; J. Wiley &Sons, Inc.: New York, 1944.

(8) Kauzmann, W. In Quantum chemistry: An introduction; Academic Press: NewYork, 1957.

(9) Crawford, T. D. Ab initio calculation of molecular chiroptical properties. Theor.Chem. Acc. 2006, 115, 227-245.

(10) Polavarapu, P. L. Renaissance in chiroptical spectroscopic methods for molecularstructure determination. Chem. Rec. 2007, 7, 125-136.

(11) Polavarapu, P. L. Ab initio molecular optical rotations and absoluteconfigurations. Mol. Phys. 1997, 91, 551-554.

(12) Autschbach, J. Computation of optical rotation using time-dependent densityfunctional theory. Comp. Lett. 2007, 3, 131-150.

(13) Giorgio, E.; Viglione, R. G.; Zanasi, R.; Rosini, C. Ab initio calculation ofoptical rotatory dispersion (ORD) curves: A simple and reliable approach to theassignment of the molecular absolute configuration. J. Am. Chem. Soc. 2004, 126, 12968-12976.

(14) Haesler, J.; Schindelholz, I.; Riguet, E.; Bochet, C. G.; Hug, W. Absoluteconfiguration of chirally deuterated neopentane. Nature 2007, 446, 526-529.

(15) Barron, L. D. In Molecular light scattering and Raman optical activity;Cambridge University Press: Cambridge, 2004.

(16) Barron, L. D.; Buckingham, A. D. Rayleigh and Raman Scattering fromOptically Active Molecules. Mol. Phys. 1971, 20, 1111-1119.

Page 89: ABSTRACT - Duke University

89

(17) Zuber, G.; Goldsmith, M. R.; Beratan, D. N.; Wipf, P. Towards Raman opticalactivity calculations of large molecules. ChemPhysChem 2005, 6, 595-597.

(18) Spencer, K. M.; Freedman, T. B.; Nafie, L. A. Scattered Circular-PolarizationRaman Optical-Activity. Chem. Phys. Lett. 1988, 149, 367-374.

(19) Hug, W. Visualizing Raman and Raman optical activity generation in polyatomicmolecules. Chem. Phys. 2001, 264, 53-69.

(20) Berova, N.; Nakanishi, K.; Woody, R. W. In Circular dichroism: principles andapplication; Wiley-VCH: New York, 2000.

(21) Dale, J. A.; Mosher, H. S. Nuclear magnetic-resonance enantiomer reagents -configurational correlations via nuclear magnetic-resonance chemical-shifts ofdiastereomeric mandelate, O-methylmandelate, and alpha-methoxy-alpha-trifluoromethylphenylacetate (Mtpa) esters. J. Am. Chem. Soc. 1973, 95, 512-519.

(22) Eliel, E. L.; Wilen, S. H.; Mander, L. N. In Stereochemistry of organiccompunds; Wiley-VCH: New York, 1994.

(23) Morimoto, Y.; Yata, H.; Nishikawa, Y. Assignment of the absolute configurationof the marine pentacyclic polyether (+)-enshuol by total synthesis. Angew. Chem. Int. Ed.2007, 46, 6481-6484.

(24) Nicolaou, K. C.; Snyder, S. A. Chasing molecules that were never there:Misassigned natural products and the role of chemical synthesis in modern structureelucidation. Angew. Chem. Int. Ed. 2005, 44, 1012-1044.

(25) Kondru, R. K.; Wipf, P.; Beratan, D. N. Theory-assisted determination ofabsolute stereochemistry for complex natural products via computation of molar rotationangles. J. Am. Chem. Soc. 1998, 120, 2204-2205.

(26) Wipf, P.; Lim, S. Total synthesis and structural studies of the antiviral marinenatural product hennoxazole A. Chimia 1996, 50, 157-167.

(27) Ribe, S.; Kondru, R. K.; Beratan, D. N.; Wipf, P. Optical rotation computation,total synthesis, and stereochemistry assignment of the marine natural product pitiamideA. J. Am. Chem. Soc. 2000, 122, 4608-4617.

(28) Zuber, G.; Goldsmith, M. R.; Hopkins, T. D.; Beratan, D. N.; Wipf, P.Systematic assignment of the configuration of flexible natural products by spectroscopicand computational methods: The bistramide C analysis. Org. Lett. 2005, 7, 5269-5272.

(29) Polavarapu, P. L. Why is it important to simultaneously use more than onechiroptical spectroscopic method for determining the structures of chiral molecules?Chirality 2008, 20, 664-672.

(30) Zuber, G.; Goldsmith, M. R.; Beratan, D. N.; Wipf, P. Assignment of theabsolute configuration of [n]-ladderanes by TD-DFT optical rotation calculations.Chirality 2005, 17, 507-510.

(31) Mascitti, V.; Corey, E. J. Enantioselective synthesis of pentacycloanammoxicacid. J. Am. Chem. Soc. 2006, 128, 3118-3119.

Page 90: ABSTRACT - Duke University

90

(32) Kondru, R. K.; Wipf, P.; Beratan, D. N. Atomic contributions to the opticalrotation angle as a quantitative probe of molecular chirality. Science 1998, 282, 2247-2250.

(33) Kondru, R. K.; Wipf, P.; Beratan, D. N. Structural and conformationaldependence of optical rotation angles. J. Phys. Chem. A 1999, 103, 6603-6611.

(34) Wiberg, K. B.; Wang, Y. G.; Vaccaro, P. H.; Cheeseman, J. R.; Luderer, M. R.Conformational effects on optical rotation. 2-substituted butanes. J. Phys. Chem. A 2005,109, 3405-3410.

(35) Kondru, R. K.; Lim, S.; Wipf, P.; Beratan, D. N. Synthetic and modelcomputational studies of molar rotation additivity for interacting chiral centers: Areinvestigation of van't Hoff's principle. Chirality 1997, 9, 469-477.

(36) Mukhopadhyay, P.; Zuber, G.; Beratan, D. N. Characterizing aqueous solutionconformation of a peptide backbone using Raman optical activity computations. Biophys.J. 2008.

(37) Weise, C. F.; Weisshaar, J. C. Conformational analysis of alanine dipeptide fromdipolar couplings in a water-based liquid crystal. J. Phys. Chem. A 2003, 107, 3265-3277.

(38) Poon, C. D.; Samulski, E. T.; Weise, C. F.; Weisshaar, J. C. Do bridging watermolecules dictate the structure of a model dipeptide in aqueous solution? J. Am. Chem.Soc. 2000, 122, 5642-5643.

(39) Mehta, M. A.; Fry, E. A.; Eddy, M. T.; Dedeo, M. T.; Anagnost, A. E.; Long, J.R. Structure of the alanine dipeptide in condensed phases determined by C-13 NMR. J.Phys. Chem. B 2004, 108, 2777-2780.

(40) Kim, Y. S.; Wang, J. P.; Hochstrasser, R. M. Two-dimensional infraredspectroscopy of the alanine dipeptide in aqueous solution. J. Phys. Chem. B 2005, 109,7511-7521.

(41) Takekiyo, T.; Imai, T.; Kato, M.; Taniguchi, Y. Temperature and pressure effectson conformational equilibria of alanine dipeptide in aqueous solution. Biopolymers 2004,73, 283-290.

(42) Barron, L. D.; Blanch, E. W.; Hecht, L. In Adv. Protein Chem., 2002; Vol. 62; pp51-90.

(43) Kondru, R. K.; Beratan, D. N.; Friestad, G. K.; Smith, A. B.; Wipf, P. Chiralaction at a distance: Remote substituent effects on the optical activity of calyculins A andB. Org. Lett. 2000, 2, 1509-1512.

(44) Wilson, S. M.; Wiberg, K. B.; Cheeseman, J. R.; Frisch, M. J.; Vaccaro, P. H.Nonresonant optical activity of isolated organic molecules. J. Phys. Chem. A 2005, 109,11752-11764.

(45) Kumata, Y.; Furukawa, J.; Fueno, T. The effect of solvents on the optical rotationof propylene oxide. Bull. Chem. Soc. Jpn. 1970, 43, 3920-3921.

Page 91: ABSTRACT - Duke University

91

(46) Mukhopadhyay, P.; Zuber, G.; Goldsmith, M. R.; Wipf, P.; Beratan, D. N.Solvent effect on optical rotation: A case study of methyloxirane in water.Chemphyschem 2006, 7, 2483-2486.

(47) Mukhopadhyay, P.; Zuber, G.; Wipf, P.; Beratan, D. N. Contiribution of asolute's chiral solvent imprint to optical rotation. Angew. Chem. Int. Ed. 2007, 46, 6450-6452.

(48) Losada, M.; Nguyen, P.; Xu, Y. J. Solvation of propylene oxide in water:Vibrational circular dichroism, optical rotation, and computer simulation studies. J. Phys.Chem. A 2008, 112, 5621-5627.

(49) Goldsmith, M. R.; Jayasuriya, N.; Beratan, D. N.; Wipf, P. Optical rotation ofnoncovalent aggregates. J. Am. Chem. Soc. 2003, 125, 15696-15697.

(50) Fidler, J.; Rodger, P. M.; Rodger, A. Chiral Solvent Structure around ChiralMolecules - Experimental and Theoretical-Study. J. Am. Chem. Soc. 1994, 116, 7266-7273.

(51) Neugebauer, J. Induced chirality in achiral media - How theory unravelsmysterious solvent effects. Angew. Chem. Int. Ed. 2007, 46, 7738-7740.

(52) Wang, S. H.; Canna, N. M. A molecular dynamics study of chirality transfer: Theimpact of a chiral solute on an achiral solvent. J. Chem. Phys. 2008, 129.

(53) Bonnemann, H.; Braun, G. A. Enantioselective hydrogenations on platinumcolloids. Angew. Chem. Int. Ed. 1996, 35, 1992-1995.

(54) Bieri, M.; Gautier, C.; Burgi, T. Probing chiral interfaces by infraredspectroscopic methods. PCCP 2007, 9, 671-685.

(55) Dressler, D. H.; Mastai, Y. Chiral crystallization of glutamic acid on selfassembled films of cysteine. Chirality 2007, 19, 358-365.

(56) Hu, X. B.; An, Q.; Li, G. T.; Tao, S. Y.; Liu, B. Imprinted photonic polymers forchiral recognition. Angew. Chem. Int. Ed. 2006, 45, 8145-8148.

(57) Garzon, I. L.; Reyes-Nava, J. A.; Rodriguez-Hernandez, J. I.; Sigal, I.; Beltran,M. R.; Michaelian, K. Chirality in bare and passivated gold nanoclusters. Phys. Rev. B2002, 66.

(58) Goldsmith, M. R.; George, C. B.; Zuber, G.; Naaman, R.; Waldeck, D. H.; Wipf,P.; Beratan, D. N. The chiroptical signature of achiral metal clusters induced bydissymmetric adsorbates. PCCP 2006, 8, 63-67.

(59) Gautier, C.; Burgi, T. Chiral inversion of gold nanoparticles. J. Am. Chem. Soc.2008, 130, 7077-7084.

(60) Hassey, R.; Swain, E. J.; Hammer, N. I.; Venkataraman, D.; Barnes, M. D.Probing the chiroptical response of a single molecule. Science 2006, 314, 1437-1439.

(61) Muller, T.; Wiberg, K. B.; Vaccaro, P. H. Cavity ring-down polarimetry (CRDP):A new scheme for probing circular birefringence and circular dichroism in the gas phase.J. Phys. Chem. A 2000, 104, 5959-5968.

Page 92: ABSTRACT - Duke University

92

(62) Grimme, S. Calculation of frequency dependent optical rotation using densityfunctional response theory. Chem. Phys. Lett. 2001, 339, 380-388.

(63) Stephens, P. J.; Devlin, F. J.; Cheeseman, J. R.; Frisch, M. J. Calculation ofoptical rotation using density functional theory. J. Phys. Chem. A 2001, 105, 5356-5371.

(64) Stephens, P. J.; McCann, D. M.; Cheeseman, J. R.; Frisch, M. J. Determinationof absolute configurations of chiral molecules using ab initio time-dependent densityfunctional theory calculations of optical rotation: How reliable are absoluteconfigurations obtained for molecules with small rotations? Chirality 2005, 17, S52-S64.

(65) Pecul, M.; Ruud, K. Ab initio calculation of vibrational Raman optical activity.Int. J. Quantum Chem. 2005, 104, 816-829.

(66) Ruud, K.; Taylor, P. R.; Astrand, P. O. Zero-point vibrational effects on opticalrotation. Chem. Phys. Lett. 2001, 337, 217-223.

(67) Mort, B. C.; Autschbach, J. Magnitude of zero-point vibrational corrections tooptical rotation in rigid organic molecules: A time-dependent density functional study. J.Phys. Chem. A 2005, 109, 8617-8623.

(68) Kongsted, J.; Pedersen, T. B.; Strange, M.; Osted, A.; Hansen, A. E.; Mikkelsen,K. V.; Pawlowski, F.; Jorgensen, P.; Hattig, C. Coupled cluster calculations of the opticalrotation of S-propylene oxide in gas phase and solution. Chem. Phys. Lett. 2005, 401,385-392.

(69) Klamt, A.; Schuurmann, G. Cosmo - a New Approach to Dielectric Screening inSolvents with Explicit Expressions for the Screening Energy and Its Gradient. J. Chem.Soc. Perkin Trans. 1993, 799-805.

(70) Ruud, K.; Zanasi, R. The importance of molecular vibrations: The sign change ofthe optical rotation of methyloxirane. Angew. Chem. Int. Ed. 2005, 44, 3594-3596.

(71) Flyvbjerg, H.; Petersen, H. G. Error-Estimates on Averages of Correlated Data. J.Chem. Phys. 1989, 91, 461-466.

(72) Gottarelli, G.; Osipov, M. A.; Spada, G. P. A Study of Solvent Effect on theOptical-Rotation of Chiral Biaryls. J. Phys. Chem. 1991, 95, 3879-3884.

(73) Craig, D. P.; Thirunamachandran, T. Change in Chiroptical Properties Caused byIntermolecular Coupling. Proc. R. Soc. Lond. A 1987, 410, 337-351.

(74) Stephens, P. J.; Devlin, F. J.; Cheeseman, J. R.; Frisch, M. J.; Bortolini, O.;Besse, P. Determination of absolute configuration using ab initio calculation of opticalrotation. Chirality 2003, 15, S57-S64.

(75) Su, Z.; Borho, N.; Xu, Y. J. Chiral self-recognition: Direct spectroscopicdetection of the homochiral and heterochiral dimers of propylene oxide in the gas phase.J. Am. Chem. Soc. 2006, 128, 17126-17131.

(76) Su, Z.; Xu, Y. Hydration of a chiral molecule: The propylene Oxide center dotcenter dot center dot(Water)(2) cluster in the gas phase. Angew. Chem. Int. Ed. 2007, 46,6163-6166.

Page 93: ABSTRACT - Duke University

93

(77) Lindahl, E.; Hess, B.; van der Spoel, D. GROMACS 3.0: a package for molecularsimulation and trajectory analysis. J. Mol. Model. 2001, 7, 306-317.

(78) Hockney, R. W.; Goel, S. P.; Eastwood, J. W. Quiet High-Resolution ComputerModels of a Plasma. J. Comput. Phys. 1974, 14, 148-158.

(79) Miyamoto, S.; Kollman, P. A. Settle - an Analytical Version of the Shake andRattle Algorithm for Rigid Water Models. J. Comput. Chem. 1992, 13, 952-962.

(80) Hess, B.; Bekker, H.; Berendsen, H. J. C.; Fraaije, J. LINCS: A linear constraintsolver for molecular simulations. J. Comput. Chem. 1997, 18, 1463-1472.

(81) Berendsen, H. J. C.; Postma, J. P. M.; Vangunsteren, W. F.; Dinola, A.; Haak, J.R. Molecular-Dynamics with Coupling to an External Bath. J. Chem. Phys. 1984, 81,3684-3690.

(82) Hermans, J.; Berendsen, H. J. C.; Vangunsteren, W. F.; Postma, J. P. M. AConsistent Empirical Potential for Water-Protein Interactions. Biopolymers 1984, 23,1513-1518.

(83) Jorgensen, W. L.; Maxwell, D. S.; TiradoRives, J. Development and testing ofthe OPLS all-atom force field on conformational energetics and properties of organicliquids. J. Am. Chem. Soc. 1996, 118, 11225-11236.

(84) Besler, B. H.; Merz, K. M.; Kollman, P. A. Atomic Charges Derived fromSemiempirical Methods. J. Comput. Chem. 1990, 11, 431-439.

(85) Ahlrichs, R.; Bär, M.; Baron, H.-P.; Bauernschmitt, R.; Böcker, S.; Ehrig, M.;Eichkorn, K.; Elliott, S.; Furche, F.; Haase, F. e. a., 2002.

(86) Fidler, J.; Rodger, P. M.; Rodger, A. Circular-Dichroism as a Probe of ChiralSolvent Structure around Chiral Molecules. J. Chem. Soc. Perkin Trans. 1993, 235-241.

(87) Morimoto, T.; Uno, H.; Furuta, H. Benzene ring trimer interactions modulatesupramolecular structures. Angew. Chem. Int. Ed. 2007, 46, 3672-3675.

(88) Jorgensen, W. L.; Tirado-Rives, J. Molecular modeling of organic andbiomolecular systems using BOSS and MCPRO. J. Comput. Chem. 2005, 26, 1689-1700.

(89) Eichkorn, K.; Treutler, O.; Ohm, H.; Haser, M.; Ahlrichs, R. Auxiliary Basis-Sets to Approximate Coulomb Potentials. Chem. Phys. Lett. 1995, 240, 283-289.

(90) Grimme, S.; Furche, F.; Ahlrichs, R. An improved method for density functionalcalculations of the frequency-dependent optical rotation. Chem. Phys. Lett. 2002, 361,321-328.

(91) Dewar, M. J. S.; Zoebisch, E. G.; Healy, E. F.; Stewart, J. J. P. The Developmentand Use of Quantum-Mechanical Molecular-Models .76. Am1 - a New General-PurposeQuantum-Mechanical Molecular-Model. J. Am. Chem. Soc. 1985, 107, 3902-3909.

(92) Storer, J. W.; Giesen, D. J.; Cramer, C. J.; Truhlar, D. G. Class-Iv Charge Models- a New Semiempirical Approach in Quantum-Chemistry. J. Comput.-Aided Mol. Des.1995, 9, 87-110.

Page 94: ABSTRACT - Duke University

94

(93) Chen, K.; Liu, Z.; Zhou, C.; Bracken, W. C.; Kallenbach, N. R. Spin relaxationenhancement confirms dominance of extended conformations in short alanine peptides.Angew. Chem. Int. Ed. 2007, 46, 9036-9039.

(94) Shi, Z. S.; Chen, K.; Liu, Z. G.; Kallenbach, N. R. Conformation of the backbonein unfolded proteins. Chem. Rev. 2006, 106, 1877-1897.

(95) Schweitzer-Stenner, R.; Measey, T. J. The alanine-rich XAO peptide adopts aheterogeneous population, including turn-like and polyproline II conformations. Proc.Natl. Acad. Sci. U.S.A. 2007, 104, 6649-6654.

(96) Makowska, J.; Rodziewicz-Motowidlo, S.; Baginska, K.; Makowski, M.; Vila, J.A.; Liwo, A.; Chmurzynski, L.; Scheraga, H. A. Further evidence for the absence ofpolyproline II stretch in the XAO peptide. Biophys. J. 2007, 92, 2904-2917.

(97) Makowska, J.; Rodziewicz-Motowidlo, S.; Baginska, K.; Vila, J. A.; Liwo, A.;Chmurzynski, L.; Scheraga, H. A. Polyproline II conformation is one of many localconformational states and is not an overall conformation of unfolded peptides andproteins. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 1744-1749.

(98) Schweitzer-Stenner, R.; Measey, T.; Kakalis, L.; Jordan, F.; Pizzanelli, S.; Forte,C.; Griebenow, K. Conformations of alanine-based peptides in water probed by FTIR,Raman, vibrational circular dichroism, electronic circular dichroism, and NMRspectroscopy. Biochemistry 2007, 46, 1587-1596.

(99) McColl, I. H.; Blanch, E. W.; Hecht, L.; Kallenbach, N. R.; Barron, L. D.Vibrational Raman optical activity characterization of poly(L-proline) II helix in alanineoligopeptides. J. Am. Chem. Soc. 2004, 126, 5076-5077.

(100) Gnanakaran, S.; Garcia, A. E. Validation of an all-atom protein force field: Fromdipeptides to larger peptides. J. Phys. Chem. B 2003, 107, 12555-12557.

(101) Graf, J.; Nguyen, P. H.; Stock, G.; Schwalbe, H. Structure and dynamics of thehomologous series of alanine peptides: A joint molecular dynamics/NMR study. J. Am.Chem. Soc. 2007, 129, 1179-1189.

(102) Best, R. B.; Buchete, N. V.; Hummer, G. Are current molecular dynamics forcefields too helical? Biophys. J. 2008, 95, L7-L9.

(103) Zhu, F.; Kapitan, J.; Tranter, G. E.; Pudney, P. D. A.; Isaacs, N. W.; Hecht, L.;Barron, L. D. Residual structure in disordered peptides and unfolded proteins frommultivariate analysis and ab initio simulation of Raman optical activity data. Proteins:Struct. Funct. Bioinfo. 2008, 70, 823-833.

(104) Barron, L. D.; Hecht, L.; Wilson, G. The lubricant of life: A proposal that solventwater promotes extremely fast conformational fluctuations in mobile heteropolypeptidestructure. Biochemistry 1997, 36, 13143-13147.

(105) Nafie, L. A. Infrared and Raman vibrational optical activity: Theoretical andexperimental aspects. Annu. Rev. Phys. Chem. 1997, 48, 357-386.

(106) Reiher, M.; Liegeois, V.; Ruud, K. Basis set and density functional dependenceof vibrational Raman optical activity calculations. J. Phys. Chem. A 2005, 109, 7567-7574.

Page 95: ABSTRACT - Duke University

95

(107) Zuber, G.; Hug, W. Rarefied basis sets for the calculation of optical tensors. 1.The importance of gradients on hydrogen atoms for the Raman scattering tensor. J. Phys.Chem. A 2004, 108, 2108-2118.

(108) Kapitan, J.; Baumruk, V.; Bour, P. Demonstration of the ring conformation inpolyproline by the Raman optical activity. J. Am. Chem. Soc. 2006, 128, 2438-2443.

(109) Kapitan, J.; Baumruk, V.; Kopecky, V.; Bour, P. Conformational flexibility of L-alanine Zwitterion determines shapes of Raman and Raman optical activity spectralbands. J. Phys. Chem. A 2006, 110, 4689-4696.

(110) Liegeois, V.; Quinet, O.; Champagne, B. Vibrational Raman optical activity as amean for revealing the helicity of oligosilanes: A quantum chemical investigation. J.Chem. Phys. 2005, 122.

(111) Liegeois, V.; Quinet, O.; Champagne, B. Investigation of polyethylene helicalconformations: Theoretical study by vibrational Raman optical activity. Int. J. QuantumChem. 2006, 106, 3097-3107.

(112) Pecul, M.; Larnparska, E.; Cappelli, C.; Frediani, L.; Ruud, K. Solvent effects onRaman optical activity spectra calculated using the polarizable continuum model. J. Phys.Chem. A 2006, 110, 2807-2815.

(113) Herrmann, C.; Ruud, K.; Reiher, M. Can Raman optical activity separate axialfrom local chirality? A theoretical study of helical deca-alanine. Chemphyschem 2006, 7,2189-2196.

(114) Madison, V.; Kopple, K. D. Solvent-Dependent Conformational Distributions ofSome Dipeptides. J. Am. Chem. Soc. 1980, 102, 4855-4863.

(115) Tran, H. T.; Wang, X. L.; Pappu, R. V. Reconciling observations of sequence-specific conformational propensities with the generic polymeric behavior of denaturedproteins. Biochemistry 2005, 44, 11369-11380.

(116) Drozdov, A. N.; Grossfield, A.; Pappu, R. V. Role of solvent in determiningconformational preferences of alanine dipeptide in water. J. Am. Chem. Soc. 2004, 126,2574-2581.

(117) Gnanakaran, S.; Hochstrasser, R. M. Conformational preferences and vibrationalfrequency distributions of short peptides in relation to multidimensional infraredspectroscopy. J. Am. Chem. Soc. 2001, 123, 12886-12898.

(118) Marsili, S.; Barducci, A.; Chelli, R.; Procacci, P.; Schettino, V. Self-healingumbrella sampling: A non-equilibrium approach for quantitative free energy calculations.J. Phys. Chem. B 2006, 110, 14011-14013.

(119) Laio, A.; Parrinello, M. Escaping free-energy minima. Proc. Natl. Acad. Sci.U.S.A. 2002, 99, 12562-12566.

(120) Smith, P. E. The alanine dipeptide free energy surface in solution. J. Chem. Phys.1999, 111, 5568-5579.

Page 96: ABSTRACT - Duke University

96

(121) Ensing, B.; De Vivo, M.; Liu, Z. W.; Moore, P.; Klein, M. L. Metadynamics as atool for exploring free energy landscapes of chemical reactions. Acc. Chem. Res. 2006,39, 73-81.

(122) Rosso, L.; Abrams, J. B.; Tuckerman, M. E. Mapping the backbone dihedral free-energy surfaces in small peptides in solution using adiabatic free-energy dynamics. J.Phys. Chem. A 2005, 109, 4162-4167.

(123) Duan, Y.; Wu, C.; Chowdhury, S.; Lee, M. C.; Xiong, G. M.; Zhang, W.; Yang,R.; Cieplak, P.; Luo, R.; Lee, T.; Caldwell, J.; Wang, J. M.; Kollman, P. A point-chargeforce field for molecular mechanics simulations of proteins based on condensed-phasequantum mechanical calculations. J. Comput. Chem. 2003, 24, 1999-2012.

(124) Han, W. G.; Jalkanen, K. J.; Elstner, M.; Suhai, S. Theoretical study of aqueousN-acetyl-L-alanine N '-methylamide: Structures and Raman, VCD, and ROA spectra. J.Phys. Chem. B 1998, 102, 2587-2602.

(125) Jalkanen, K. J.; Degtyarenko, I. M.; Nieminen, R. M.; Cao, X.; Nafie, L. A.; Zhu,F.; Barron, L. D. Role of hydration in determining the structure and vibrational spectra ofL-alanine and N-acetyl L-alanine N'-methylamide in aqueous solution: a combinedtheoretical and experimental approach. Theor. Chem. Acc. 2008, 119, 191-210.

(126) Deng, Z.; Polavarapu, P. L.; Ford, S. J.; Hecht, L.; Barron, L. D.; Ewig, C. S.;Jalkanen, K. Solution-phase conformations of N-acetyl-N'-methyl-L-alaninamide fromvibrational Raman optical activity. J. Phys. Chem. 1996, 100, 2025-2034.

(127) Lovell, S. C.; Davis, I. W.; Adrendall, W. B.; de Bakker, P. I. W.; Word, J. M.;Prisant, M. G.; Richardson, J. S.; Richardson, D. C. Structure validation by C alphageometry: phi,psi and C beta deviation. Proteins: Struct. Funct. Genet. 2003, 50, 437-450.

(128) Zhu, F. J.; Isaacs, N. W.; Hecht, L.; Barron, L. D. Raman optical activity: A toolfor protein structure analysis. Structure 2005, 13, 1409-1419.

(129) Kapitan, J.; Baumruk, V.; Pohl, R.; Bour, P. Proline zwitterion dynamics insolution, glass and crystalline state. J. Am. Chem. Soc. 2006, 128, 13451-13462.

(130) Zuber, G.; Hug, W. Computational interpretation of vibrational optical activity:The ROA spectra of (4S)-4-methylisochromane and the (4S)-isomers of Galaxolide((R)).Helv. Chim. Acta 2004, 87, 2208-2234.

(131) Chellgren, B. W.; Creamer, T. P. Effects of H2O and D2O polyproline lane IIhelical structure. J. Am. Chem. Soc. 2004, 126, 14734-14735.

(132) Herrmann, C.; Ruud, K.; Reiher, M. Importance of backbone angles versusamino acid configurations in peptide vibrational Raman optical activity spectra. Chem.Phys. 2007, 343, 200-209.

(133) Liegeois, V.; Ruud, K.; Champagne, B. An analytical derivative procedure forthe calculation of vibrational Raman optical activity spectra. J. Chem. Phys. 2007, 127.

(134) Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R. W.; Klein, M. L.Comparison of Simple Potential Functions for Simulating Liquid Water. J. Chem. Phys.1983, 79, 926-935.

Page 97: ABSTRACT - Duke University

97

(135) Allen, M. P.; Tildesley, D. J. In Computer simulations of liquids; OxfordUniversity Press: Oxford, 1987.

(136) Luber, S.; Herrmann, C.; Reiher, M. Relevance of the Electric-Dipole-Electric-Quadrupole Contribution to Raman Optical Activity Spectra. J. Phys. Chem. B 2008, 112,2218-2232.

(137) Barron, L. D.; Escribano, J. R. Stokes-Antistokes Asymmetry in Natural RamanOptical-Activity. Chem. Phys. 1985, 98, 437-446.

(138) Eichkorn, K.; Weigend, F.; Treutler, O.; Ahlrichs, R. Auxiliary basis sets formain row atoms and transition metals and their use to approximate Coulomb potentials.Theor. Chem. Acc. 1997, 97, 119-124.

(139) Zuber, G.; Wipf, P.; Beratan, D. N. Exploring the optical activity tensor byanisotropic Rayleigh optical activity scattering. ChemPhysChem 2008, 9, 265-271.

(140) Losada, M.; Xu, Y. J. Chirality transfer through hydrogen-bonding: Experimentaland ab initio analyses of vibrational circular dichroism spectra of methyl lactate in water.PCCP 2007, 9, 3127-3135.

(141) Losada, M.; Tran, H.; Xua, Y. J. Lactic acid in solution: Investigations of lacticacid self-aggregation and hydrogen bonding interactions with water and methanol usingvibrational absorption and vibrational circular dichroism spectroscopies. J. Chem. Phys.2008, 128.

(142) Glattli, A.; Daura, X.; Seebach, D.; van Gunsteren, W. F. Can one derive theconformational preference of a beta-peptide from its CD spectrum? J. Am. Chem. Soc.2002, 124, 12972-12978.

(143) Liu, J.; Wang, D.; Zheng, Q.; Lu, M.; Arora, P. S. Atomic structure of a shortalpha-helix stabilized by a main chain hydrogen-bond surrogate. J. Am. Chem. Soc. 2008,130, 4334-4337.

(144) Kumin, M.; Sonntag, L. S.; Wennemers, H. Azidoproline containing helices:Stabilization of the polyproline II structure by a functionalizable group. J. Am. Chem.Soc. 2007, 129, 466-467.

(145) Wright, P. E.; Dyson, H. J. Intrinsically unstructured proteins: Re-assessing theprotein structure-function paradigm. J. Mol. Bio. 1999, 293, 321-331.

(146) Iakoucheva, L. M.; Brown, C. J.; Lawson, J. D.; Obradovic, Z.; Dunker, A. K.Intrinsic disorder in cell-signaling and cancer-associated proteins. J. Mol. Biol. 2002,323, 573-584.

Page 98: ABSTRACT - Duke University

98

Biography

Parag Mukhopadhyay earned a B.Sc. with honors in chemistry from the University of

Delhi, India and M.Sc. degrees in chemistry and biological sciences from the Indian Institute of

Technology, Delhi, India, and the University of Calgary, Canada, respectively. He conducted his

doctoral research in theoretical and experimental probes of molecular chirality at Duke

University, and published the following articles:

I. P. Mukhopadhyay, P. Wipf, D.N. Beratan. “Optical signatures of molecular

dissymmetry: Combining theory with experiments to address stereochemical

puzzles”, to be submitted, Accounts of Chemical Research.

II. P. Mukhopadhyay, G. Zuber, D.N. Beratan. “Characterizing aqueous solution

conformations of a peptide backbone using Raman optical activity computation”,

Biophysical Journal, 2008. In press.

III. P. Mukhopadhyay, G. Zuber, P. Wipf, D.N. Beratan. “Contribution of a solute’s

chiral solvent imprint to optical rotation”, Angewandte Chemie International

Edition, 46:6450-6452, 2007. The work was also highlighted in Science (2007,

317:725), Chemical and Engineering News (6 August, 2007) and Angewandte

Chemie, International Edition (2007, 46:7738-7740).

IV. P. Mukhopadhyay, G. Zuber, M.R. Goldsmith, P. Wipf, D.N. Beratan. “Solvent

effects on optical rotation: A case study of methyloxirane in water”,

ChemPhysChem, 7:2483-2486, 2006.