Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 May 2.
Published in final edited form as: J Phys Chem B. 2019 Jun 27;123(27):5769–5781. doi: 10.1021/acs.jpcb.9b04029

Chiral Inversion of Amino Acids in Anti-parallel β-sheets at Interfaces Probed by Vibrational Sum Frequency Generation Spectroscopy

Ethan A Perets a,, Pablo E Videla a,b,, Elsa C Y Yan a,*, Victor S Batista a,b,*
PMCID: PMC9059514  NIHMSID: NIHMS1797907  PMID: 31194546

Abstract

Parallel study of protein variants with all (L-), all (D-) or mixed (L-)/(D-) amino acids can be used to assess how backbone architecture versus sidechain identity determines protein structure. Here, we investigate the secondary structure and sidechain orientation dynamics of the anti-parallel β-sheet peptide LK7β (Ac-Leu-Lys-Leu-Lys-Leu-Lys-Leu-NH2) composed of all (L-), all (D-), or alternating (L-Leu)/(D-Lys) amino acids. Using interface-selective vibrational sum frequency generation spectroscopy (VSFG), we observe that the alternating (L-)/(D-) peptide lacks a resonant C-H stretching mode compared to the (L-) and (D-) variants, and does not form anti-parallel β-sheets. We rationalize our observations based on density functional theory calculations and molecular dynamics (MD) simulations of LK7β at the air-water interface. Irrespective of the handedness of the amino acids, leucine sidechains prefer to orient toward the hydrophobic air phase while lysine sidechains prefer the hydrophilic water phase. These preferences dictate the backbone configuration of LK7β and thereby the folding of the peptide. Our MD simulations show that the preferred sidechain orientations can force the backbone of a single strand of (L-) LK7β at the air-water interface to adopt β-sheet Ramachandran angles. However, denaturation of the β-sheets at pH = 2 results in negligible chiral VSFG amide I response. The combined computational and experimental results lend critical support to theory that chiral VSFG response requires macroscopic chirality, such as in β-sheets. Our results can guide expectations about the VSFG optical responses of proteins, and should improve understanding of how amino acid chirality modulates the structure and function of natural and de novo proteins at biological interfaces.

Graphical Abstract

graphic file with name nihms-1797907-f0001.jpg

Introduction

The enantiomeric enrichment of (L-) chirality amino acids placed significant restrictions on the chemical diversity of proteins in the pre-biotic world. However, (L-) and (D-) amino acids are occasionally incorporated together in natural systems. For example, the antibiotic peptide gramicidin A consists of fifteen alternating (L-) and (D-) amino acids. The peptide adopts the uncommon β-helix-type secondary structure with all sidechains on one face of the strand, allowing gramicidin A to function as a membrane cation channel.12 The incorporation of both (L-) and (D-) amino acids in natural proteins requires a complex biosynthetic machinery acting along non-ribosomal pathways.34 This raises questions concerning the unique biophysical properties of heterochirality that would justify expanding “Nature’s alphabet” to include (D-) amino acids.5

(D-) amino acids have been used in the design and synthesis of unnatural proteins.610 Some of the resulting proteins with (D-) amino acids show promise as anti-viral and anti-cancer therapeutic agents, since those unnatural proteins are more resistant to proteolysis, may exchange their chiral recognition in biomolecular interactions, and modulate the biophysical properties of cellular structures.9, 11 For example, the antimicrobial peptide sapesin B composed of only (D-) leucine and lysine has higher antimicrobial activity than all (L-) sapesin B.12 The improved therapeutic potential has been attributed to distinct interactions between (D-) sapesin B and the cell wall of Staphylococcus bacteria. At the molecular level, recent computational work employing all-atom molecular dynamics has systematically studied the impact of mixed (L-) and (D-) chirality on a-helices.13 The results showed that the (L-) to (D-) chiral inversion of even a single amino acid can locally break α-helix secondary structure. An outstanding challenge is to understand the perturbations induced by chiral inversion leading to disruption of protein secondary structure.

Vibrational sum frequency generation spectroscopy (VSFG) is a second-order (χ(2)) vibrational spectroscopy that can distinguish between protein secondary structures.1423 VSFG is a three-wave mixing process that irradiates the sample with pulses of visible and IR light with frequencies ωvis and ωIR, respectively, and detects the vibrationally-resonant sum frequency response at frequency ωSFG = ωvis + ωIR. Due to non-linear effects of the three-wave mixing process, the VSFG technique is highly sensitive and selective for vibrational probes in anisotropic environments, such as interfaces, and suppresses the background signal from the bulk.14, 18, 2428 When combined with computational modeling, VSFG can provide information on molecular orientation at interfaces.2733 The SFG, visible, and IR beams can each be polarized parallel (p) or perpendicular (s) to the incident plane, yielding eight possible polarization combinations. Comparison of SFG intensities using at least two different polarization combinations can, for example, be used to assess interfacial orientation of molecules, such as lipid tails and alkyl chains when probing the CH3 stretching bands.29, 3438

In this paper, we combine achiral and chiral VSFG spectroscopy to probe the perturbations of amino acid chiral inversions on the overall chiral secondary structure of anti-parallel β-sheets at interfaces. The experimental data is interpreted through density functional theory (DFT) calculations of the VSFG spectra and protein backbone geometry and bond orientations in molecular dynamic (MD) simulations of the anti-parallel β-sheet at the air-water interface. We find that the VSFG technique can distinguish all (L-) or all (D-) anti-parallel β-sheets from a variant that contains alternating (L-) and (D-) chirality along the length of the peptide backbone, proving the power of VSFG as a vibrational chiroptical spectroscopy. At the molecular level, we find that the presence of an amphiphilic interface induces a preferential orientation of hydrophobic and hydrophilic amino acid sidechains to be exposed to the air or water phases, respectively. These preferred orientations in the peptide sequence of LK7β confine the Ramachandran angles to β-sheet-like configuration even for a single strand of (L-) LK7β, providing an invaluable molecular system to critically examine the chiral SFG theory.3942 The mixed (L-)/(D-) chirality conforms to the preferences of sidechain orientation and thereby exerts steric constraints that prevent the peptide backbone from forming anti-parallel β-sheets. Revealing this novel aspect of how chirality and chemical properties of side chains can modulate the secondary structure and orientation of proteins at interfaces exemplifies the advantages of employing both experiments and theoretical analyses for interpreting VSFG optical response of peptides and proteins. Moreover, our findings should offer insights into de novo design of unnatural proteins with both (L-) and (D-) amino acids with potential applications for engineering interfaces of biomaterials.

Methods

Sample Preparation

The LK7β (Acetyl-Leu-Lys-Leu-Lys-Leu-Lys-Leu-NH2, MW: 896.24 g/mol) with all (L-) amino acids was synthesized by GL Biochem Ltd. (Shanghai). The LK7β with all (D-) or (L-)/(D-) amino acids (i.e., Acetyl-(L-)Leu-(D-)Lys-(L-)Leu-(D-)Lys-(L-)Leu-(D-)Lys-(L-)Leu-NH2) were synthesized by AnaSpec, Inc. (Fremont, CA, USA). The lyophilized powder samples were dissolved in deionized H2O (pH =7 or pH = 2, HCl adjusted) at 1 mM. The solutions were prepared as 100 uL aliquots, frozen in liquid nitrogen, and stored at −80 °C until use.

For VSFG experiments at the air-glass interface, plain glass microscope slides (Thermo Scientific, Portsmouth, NH, USA; Cat No. 420–004T, 101616–3) were plasma-cleaned on “high” for 3 minutes in a plasma cleaner (Harrick Plasma, Ithaca, NY, USA; PDC-32G). One aliquot of the LK7β was thawed and pipetted onto the glass slides. The samples dried overnight at room temperature, forming hydrated thin-films at the air-glass interface.

For VSFG experiments at the air-water interface, 200 μL of 1mM LK7β solution was applied with a glass syringe (Hamilton Company, Reno, NV, USA; Model No. 1705N) at the surface of 3.8 mL of deionized water, giving a final concentration of LK7β of 50 μM.

VSFG Measurements

The VSFG spectra were collected with a home-built setup previously described.43 For achiral VSFG measurements, we centered the broad-band IR beam at 3300 nm and used the ssp polarization configuration (s-polarized sum frequency, s-polarized visible, and p-polarized IR). For chiral amide I/amide II VSFG measurements, we centered the broad-band IR beam at 6100 nm and used the psp polarization configuration. Spectra taken at the air-glass interface were collected with acquisition times of 10 minutes (unless otherwise noted). Spectra taken at the air-water interface were collected with acquisition times of 20 minutes.

The reported IR frequencies are recovered by calibrating the VSFG response to a polystyrene standard (Buck Scientific, East Norwalk, CT, USA; 0.05 mm film). Background spectra were collected by shuttering the IR beam. The background spectra were then subtracted from the VSFG intensity spectra. For VSFG spectra at the air-glass and air-water interfaces, the background-subtracted spectra were normalized to the IR power envelope by dividing by the VSFG response of a GaAs crystal. The VSFG spectra were manually cleaned of cosmic ray intensities, which are unrelated to the VSFG response.

Finally, the intensity spectra were fit to a Lorentzian function that describes the resonant vibrational mode being probed and a contributing non-resonant term:

ISFG|χNR(2)+qAqωIRωq+iΓq|2 (1)

where ISFG is the intensity of the sum frequency generation, χNR(2) is the non-resonant second order susceptibility, ωIR is the frequency of the IR beam, Aq is the amplitude of the qth resonant vibrational mode, ωq is the frequency, and Гq is the half-width half maximum.

Molecular Dynamics

Molecular dynamics simulations of all (L-), all (D-) and (L-Leu)/(D-Lys) LK7β on the water/air interface were performed using the all-atoms CHARMM36/CMAP force field44 and the TIP3P water model potential.45 The force field parameters for the (D-) amino acids were taken from the residue topology ‘toppar_all36_prot_d_aminoacids.str’ that includes inverted CMAP tables from the corresponding (L-) amino acids. All simulations were performed using the NAMD software package46 in the NVT ensemble using a Langevin thermostat with a time constant of 1 ps. Hydrogen bonds were kept fixed using the SETTLE algorithm,47 which allows the use of a 2 fs time step. Periodic boundaries conditions in three dimensions were applied using a cutoff of 12 Å for the real-space Coulombic and van der Waals interactions, whereas long-range interactions were handle using the particle-mesh Ewald method (PME).

The initial LK7 polypeptides strands were built in the β-conformation by tacking backbone dihedral angles (ϕ,ψ) of (−120,113) for L-amino acids and (120,−113) for D-amino acids, using the Molefacture plugin in VMD package.48 The sidechains of the lysine residues were taken to be charged. The N-terminal and C-terminal of the polypeptide chains were capped with −NH2COCH3 (acetylated N-terminus) and −CONH2 (amidated C-terminus) groups respectively. From these polypeptide strands, β sheets were assembled in antiparallel conformation using 2 strands.

The β sheets were placed on the surface of a preequilibrated TIP3P water slab with the hydrophobic leucine residues facing the air. The water slab consists of 4855 to 4893 water molecules arranged in a ~35 Å thick slab of 70×70 Å2 surface area, with ~35 Å of vacuum in the direction perpendicular to the slab (defined as the z direction). In order to maintain charge neutrality, random waters were replaced by Cl atoms. The combined protein-water system was allowed to relax using conjugate gradient algorithm for 5000 steps to eliminate any close contact interaction between atoms, followed by an 820 ps simulation where the system was heated from 30 K to 298 K. An additional 400 ps simulation at 298 K was carried out to equilibrate the system at the target temperature. For all (L-) and all (D-) LK7β, a 20 ns production simulation was used to extract dihedral angles and angular distribution using snapshots every 1 ps. For (L-Leu)/(D-Lys) LK7β, longer simulations of 60 ns were performed. Results for the all (L-) and all (D-) LK7β correspond to averages over three independent simulations with different initial configurations, whereas results for the (L-Leu)/(D-Lys) LK7β model correspond to averages over five simulations.

Calculation of VSFG Spectra

VSFG spectra were computed for an antiparallel β-sheet model composed of two LK3 peptide, with the primary sequence Leu-Lys-Leu. This peptide represents a minimal model with the characteristics of the LK7β peptide. Geometry optimizations and harmonic frequency analysis at the DFT level were performed with the Gaussian 2009 software package,49 using the B3LYP50 hybrid functional and the 6–31G(d) basis set.51 An “ultrafine” integration grid (99 radial shells and 590 angular points per shell) was used for the frequency calculations to obtain accurate results.

Our approach for simulating SFG spectra has been previously described.17, 20, 5258 Here, we describe it only briefly. The calculations involves the determination of the second-order molecular hyperpolarizabilities βijk,q(2)αijQqμkQq where αij and μk (with i, j, k = a, b, c) are elements of the polarizability and dipole moment in the molecular frame, respectively, and Qq is the normal coordinate of the q-th vibrational mode. The hyperpolarizability βijk,q(2) is then rotated to the laboratory frame (x, y, z) and averaged over the azimuthal angle ϕ in 5° increments to obtain the second-order susceptibility χIJK,q(2)=ijkRIiRJjRKkβijk,q(2) where RIi represent elements of the ZYZ Euler rotation matrix.18 The susceptibilities χIJK,q(2) are then used to compute the effective susceptibilities, which for the ssp and psp polarization combinations take the form18, 27, 29

χssp,q(2)=Lyyzχyyz,q(2) (2)
χpsp,q(2)=Lzyxχzyx,q(2)Lxyzχxyz,q(2) (3)

where Lyyz are Fresnel factors that depend on the refractive index of the interface as well as the incident angle of the lights.18, 27, 29 The Fresnel factors used in this study are listed in Table S2. From these susceptibilities, the homodyne SFG spectra is given by

ISFG(ωIR)|χ(2)|2=|ANReiϕ+kqχq(2)ωIRωq+iΓq|2 (4)

where k is a multiplicative constant. All harmonic frequencies were scaled by 0.943 to facilitate comparisons with experiments whereas Γq and ANR were set to 7.5 cm−1 and 0, respectively.

Results

Achiral and Chiral VSFG of (L-), (D-), and (L-Leu)/(D-Lys) LK7β.

We previously demonstrated by VSFG amide I response that the (L-) LK7β peptide spontaneously forms anti-parallel β-sheets at interfaces.54 The LK7β peptide localizes to the air-water interface due to alternating hydrophobic leucine and hydrophilic lysine residues.59 Figure 1 shows the achiral (ssp) VSFG spectra of (L-), (D-) and (L-Leu)/(D-Lys) LK7β variants at the air-water interface. The (L-) LK7β shows three peaks in the C-H stretching region (Figure 1a) in agreement with the spectrum reported by Somorjai and co-workers.60 The peaks are centered at 2875 cm−1, 2913 cm−1 and 2938 cm−1. Their assignments are aided by computational studies as discussed in the next section (Computations of VSFG Spectra). The (L-) LK7β and (D-) LK7β spectra are indistinguishable (Figure 1b, green and gray). However, the achiral VSFG spectrum for (L-Leu)/(D-Lys) LK7β is distinguished by significant suppression of the peak at 2913 cm−1 (Figure 1b, blue). Suppression of this peak is also observed at the air-glass interface (Figures 2a and 2b). In contrast, linear vibrational spectroscopy such as attenuated total reflectance-Fourier transform infrared does not distinguish between the three LK7β species in the C-H stretching region (Figure S1, Section S1 in Supporting Information).

Figure 1.

Figure 1.

VSFG of (L-), (D-), and (L-Leu)/(D-Lys) LK7β at the air-water interface in the C-H stretching region. The spectra of (L-) and (D-) LK7β show identical profiles whereas the (L-Leu)/(D-Lys) variant shows strong suppression of the peak at 2913 cm−1. The spectra are measured using the ssp (achiral) polarization combination. Solid colored lines are fits to the spectra based on equation 1. Solid black lines are the component peaks of the fits.

Figure 2.

Figure 2.

VSFG of (L-), (D-), and (L-Leu)/(D-Lys) LK7β at the air-glass interface in the C-H stretching region. Similar to the air-water interface, the peak at 2911 cm−1 is suppressed in the (L-Leu)/(D-Lys) variant. The spectra in (a) and (b) are measured using the ssp (achiral) polarization combination. The spectra in (c) and (d) are measured with the psp (chiral) polarization combination. Solid colored lines are fits to the spectra based on equation 1. Solid black lines are the component peaks of the fits.

We also probed LK7β variants at the air-glass interface using chiral (psp) VSFG, which is sensitive to anisotropic orientation of local chiral moieties as well as macroscopic chirality.23, 4142 The chiral VSFG response can report on formation of higher-order chiral structures of biological macromolecules, for example, α-helix or β-sheet secondary structure of proteins and hybridization of double-helix DNA.19, 23, 61 The (L-) LK7β peptide shows at least four peaks in the C-H stretching region (Figure 2c), in agreement with our prior report of chiral VSFG spectra of LK7β at the air-water interface.62 The peak positions, centered at 2869 cm−1, 2918 cm−1, 2965 cm−1 and 2988 cm−1 were previously assigned to the CH3 symmetric stretch, CH2 asymmetric stretch, CH3 asymmetric stretch, and Cα-H stretching modes, respectively.62 The (D-) LK7β spectrum is very similar and overall indistinguishable (Figure 2d, green and gray). In contrast, chiral response for (L-Leu)/(D-Lys) LK7β was not observed (Figure 2d, blue). These observations suggest that (L-Leu)/(D-Lys) LK7β at the air-glass interface neither assumes a configuration in which local chiral moieties produce a coherent VSFG response, nor adopts a chiral secondary structure.

We have also analyzed the effect of pH by preparing (L-) LK7β at the air-glass interface at neutral pH = 7 and denaturing pH = 2. We find that the peak at 2913 cm−1 is clearly not suppressed at pH = 2 (Figure 3a). Furthermore, we measured the chiral amide I/amide II response of (L-) LK7β at the air-glass interface at both pH = 7 and pH = 2. We have previously shown that the chiral amide I response is a useful probe of the protein secondary structure at interfaces.16, 18 For LK7β, the peaks at 1563 cm−1, 1618 cm−1, and 1687 cm−1 are assigned to the amide II, amide I (B2), and amide I (B1) vibrational modes associated with anti-parallel β-sheets, respectively (Figure 3b, left).18, 54 However, the chiral amide I/amide II spectrum measured with the psp polarization is almost completely abolished at pH = 2 (Figure 3b, right). This suggests that the anti-parallel β-sheet character of the peptide is mostly lost under denaturing conditions. Altogether, these results show that the loss of anti-parallel β-sheet structure is not correlated with suppression of the peak at 2913 cm−1 as observed for the (L-Leu)/(D-Lys) LK7β spectrum (Figures 1b and 2b). Finally, Figure 3c shows the chiral amide I/amide II spectrum measured for the different LK7β variants at pH = 7. The (L-) and (D-) LK7β spectra show strong chiral response, while no VSFG signal is observed for (L-Leu)/(D-Lys) LK7β.

Figure 3.

Figure 3.

Achiral and chiral VSFG of LK7β at the air-glass interface prepared using native pH = 7 and denaturing pH = 2 conditions. (a,b) Loss of anti-parallel β-sheet structure at pH = 2 induces loss of chiral amide I/amide II signal, which is not correlated with suppression of the C-H stretching peak at 2913 cm−1. (c) (L-Leu)/(D-Lys) LK7β does not show chiral VSFG signal in the amide I/amide II region. The spectra in (a) are measured using the ssp (achiral) polarization combination. The pH = 2 spectra in (a) was scaled by 2x. The spectra in (b) and (c) are measured with the psp (chiral) polarization combination.

The suppression of the VSFG response of the resonant vibrational mode at 2913 cm−1 for (L-Leu)/(D-Lys) LK7β (Figures 1 and 2) and our finding that the suppression of this mode is not correlated with loss of anti-parallel β-sheet structure (Figure 3) prompts some interesting questions. What is the identity of the vibrational normal mode at 2913 cm−1? Why is this mode suppressed in heterochiral versus homochiral LK7β variants? How do chiral inversions inhibit or promote protein secondary structure at interfaces? To address these research questions in molecular detail, we used DFT to recapitulate the VSFG response of (L-) LK7β. We also performed MD simulations to understand the effects of chiral inversions on protein secondary structure and sidechain dynamics of (L-), (D-) and (L-Leu)/(D-Lys) LK7β at the air-water interface.

Computations of VSFG spectra

To approximate the VSFG response of (L-) LK7β and assign the peak at 2913 cm−1, we performed DFT calculations on an antiparallel β-sheet model composed of two (L-) LK3 peptides with primary sequence Ac-Leu-Lys-Leu-NH2 (Figure 4a). The model peptide represents a minimal ‘building block’ with the characteristics of (L-) LK7β amenable to ab initio calculations. While treatments based on second-order perturbation theory (VPT2)6366 or vibrational self-consistent field (VSCF)6769 methods are available to include overtones and combination bands in the description of the spectra, the dimension of the system under consideration makes such calculations infeasible. Nevertheless, we find that an anharmonic analysis comprising Fermi resonances (FRs) on single leucine and lysine residues including couplings between modes7072 is sufficient for understanding the C-H stretch spectral region (Section S2 in Supporting Information).

Figure 4.

Figure 4.

(a) Molecular representation of the antiparallel β-sheet all (L-) (LK3)2 model peptide, defined in the molecular frame a, b, c. DFT-based (b) ssp and (c) psp SFG spectra. Shaded areas correspond to symmetric CH2 (purple), symmetric CH3 (red), asymmetric CH2 (green) and asymmetric CH3 (cyan) stretching modes.

Figure 4b shows the calculated ssp VSFG spectrum of the antiparallel β-sheet model for a molecular orientation with the c-axis perpendicular to the surface plane (axis coordinates are shown in Figure 4a). This orientation exposes the hydrophobic leucine sidechains to the hydrophobic phase and the charged lysine sidechains to the hydrophilic phase. This is the dominant structure of the anti-parallel β-sheet at the air-water interface (see next section). The spectrum is characterized by a broad peak centered around 2875 cm−1 attributed to the symmetric CH3 stretching (normal mode analyses are presented in Section 3 in Supporting Information). This peak has shoulders at both sides due to symmetric and asymmetric CH2 stretching (note that a shoulder possibly appears around 2850 cm−1 in the experimental spectrum in Figure 1). Based on this analysis, the peak at 2875 cm−1 in the experimental spectra in Figures 1 and 2 can be assigned to symmetric CH3 stretch. Allowing for the contributions of FRs of the symmetric CH3 and CH2 stretching of leucine and lysine sidechains (Supporting Information) and given the strong intensity of the symmetric CH3 and CH2 peaks, from which intensity of the FR is borrowed, we assign the peak at 2913 cm−1 to FR of the symmetric CH2 stretching and the peak at 2938 cm−1 to CH3 FR. The lysine Cα-H is coupled to the CH2 FRs and possibly also contributes in the experimental spectrum. This assignment, based on DFT calculations, is in agreement with previous studies of the LK7β polypeptide as well as individual leucine and lysine residues.22, 33, 60, 7375 It should be noted that the symmetric CH3 stretching band of the capping group appears at ~2893 cm−1 and might also contribute to the spectral signals. We cannot presently discard the contribution of this capping group to the experimental spectra. However, the reasonable agreement in peak position and relative intensity between the calculated and experimental spectra, as well as previous SFG studies of LK7β and related systems such as LK14α, leucine, and lysine amino acids,22, 33, 60, 7377 suggest that the contribution of the capping to the SFG signal should be small.

In Figure 4c we present the chiral spectra for the model peptide. The simulated and experimental spectra (Figure 2c) agree very well. The spectra consist of four bands, including symmetric CH3, asymmetric CH2, asymmetric CH3 and Cα-H stretching modes.62 As noted above, the asymmetric CH2 stretching modes that appear in the region 2900–2930 cm−1 are highly coupled to the chiral Cα-H of the lysine residues. The higher frequency peaks comprise the stretching of the leucine chiral Cα-H moieties (with some possible overlap with the asymmetric CH3 stretching of the acetyl capping group). For the spectrum in Figure 4c, it is difficult to assess the effects of FRs in the chiral SFG signal. Further studies are required.

Molecular Dynamics and Orientation Analysis of (L-), (D-), and (L-Leu)/(D-Lys) LK7β.

To obtain molecular detail on the interfacial behavior of the LK7β system, we complement the VSFG experiments and calculations with MD simulations. The simulated systems consist of a water slab with (L-), (D-), or (L-Leu)/(D-Lys) LK7β placed at the air-water interface (Figure 5).

Figure 5.

Figure 5.

Typical snapshots of the molecular dynamics simulations of LK7β at the air-water interface. (a,b) Side view and top view of (L-) LK7β. (c,d) Side view and top view of (L-Leu)/(D-Lys) LK7β.

We first characterize the backbone conformation of polypeptides with different chirality by computing free energy profiles along the backbone dihedral angles ϕ and ψ. In Figure 6a we present the free energy profile for (L-) LK7β composed of two strands in an antiparallel β-sheet conformation at the air-water interface. The free energy landscape presents a well spanning the region −160°<ϕ<−80° and 100°<ψ<150°, characteristic of β-sheet conformations.78 An analysis of the free energy profile per residue (Figure S5) also corroborates the presence of β-sheet conformations. In this conformation all the hydrophilic lysine residues are placed on one side of the sheet and are solvated by the water phase; all the hydrophobic leucine residues point into the air (Figure 5a). In fact, a simulation test in which the LK7β is started with the leucine residues pointing into the water and the lysine residues into the air, shows that the system naturally evolves to invert the orientations of the residues (Figure S7).

Figure 6.

Figure 6.

Free energy (kBT units) as a function of the dihedral angles ϕ and ψ for (a) (L-) (two-strands), (b) (D-) (two-strands), (c) (L-Leu)/(D-Lys) (one-strand) and (d) (L-) (one-strand) and LK7β.

Figure 6b shows the Ramachandran free energy profile for a two-strand (D-) LK7β system. The free energy landscape obtained for this system is a mirror image of the one for the (L-) LK7β (Figure 6a), and represents the β-sheet conformational region for (D-) amino acids.78

We found that a system of two strands of (L-Leu)/(D-Lys) LK7β does not form stable anti-parallel β-sheets as the strands break apart over the course of the simulation (Figure S8, Section 6 in Supporting Information). Indeed, previous experimental and computational studies showed that inversion of chirality at a particular placement along a peptide chain leads to a consequent break in secondary structure.13, 7883 Our MD results also suggest that the (L-Leu)/(D-Lys) LK7β do not form any stable secondary structure, consistent with the VSFG silent chiral amide I/amide II spectrum (Figure 3c). Hence, for the (L-Leu)/(D-Lys) LK7β we only present results obtained with one polypeptide strand (Figure 5c,d). Figure 6c shows the free energy profile for the one-strand (L-Leu)/(D-Lys) LK7β system. The free energy landscape is richer and more complicated than for the (L-) or (D-) LK7β variants and demonstrates a more complex behavior of the polypeptide, which dynamically samples various configurations over the course of the simulation (Figure S6). The periodic alternation of chirality in the (L-Leu)/(D-Lys) LK7β system represents a huge perturbation in the “natural” behavior of the polypeptide, not allowing for the formation of local secondary structure motifs.

To further highlight the effect of chiral inversion on the dynamics and structure of peptides, we also performed simulations of a single strand of (L-) LK7β on the air-water interface. By comparing this system with the (L-Leu)/(D-Lys) LK7β and the two strands of (L-) LK7β it is possible to disentangle the effect of chiral inversion from the absence of a β-sheet. The free energy profile of the backbone for this system is presented in Figure 6d. The conformation adopted by the single strand of (L-) LK7β does not present such a complex landscape as the one adopted by the (L-Leu)/(D-Lys) LK7β (see also Figure S6), demonstrating that the inversion in chirality has a more profound effect than just disrupting the secondary structure of the peptide. Quite remarkably, while the isolated one strand of (L-) LK7β displays a richer landscape than for the system with two strands (Figure 6a), the single strand is predominantly characterized at the air-water interface by a stable backbone geometry consistent with β-strand conformation. Since it is not possible to form inter-strand hydrogen bonds, the single strand must be stabilized by the interface itself. Indeed, the hydrophobic leucine sidechains continue to point towards the air phase, while the hydrophilic lysine residues point towards the water phase (Figures S9 and S10).

The Ramachandran free energy plots presented in Figure 6 are informative with respect to the effect of chirality on the conformation of the peptide backbone but do not give information of the orientation of the sidechains with respect to the interface, which is the information gathered from achiral VSFG spectra.18 To characterize the behavior of the sidechains we computed the tilt angle (θ) of certain bonds with respect to the surface normal. A schematic representation of the representative selected bonds is presented in Figure 7.

Figure 7.

Figure 7.

(Left) Schematic representation of leucine (top) and lysine (bottom) residues with colored definition of bond vectors used to describe the orientation with respect to the surface normal. Angular distributions of selected bond vectors of leucine (a) and lysine (b) for (L-) LK7β. Angular distributions of selected bond vectors of leucine (c) and lysine (d) for (L-Leu)/(D-Lys) LK7β. Shaded areas correspond to standard deviation of the distributions. Dashed lines correspond to an isotropic distribution of the angles.

In Figure 7a we present the tilt angle distributions over all the leucine residues for two strands of (L-) LK7β. Average values and standard deviations for tilt angles obtained from these distributions are presented in Table S1. The angle θ of the Cα-Cβ bonds (blue) presents a peaked distribution with a maximum near 0° (cosθ = 1), which demonstrate the extreme preference of leucine hydrophobic sidechains to point towards the air. The tilt orientation of Cα-H bonds (black) present a broad distribution centered around θ = 90° (cosθ ≈ 0) and is consistent with the formation of anti-parallel β-sheets with hydrogen bonds between C=O---H-N lying on the plane of the water surface (Figure S11). The isopropyl group of leucine residues, characterized by the Cγ-H bond (red), presents a broad distribution centered around 120° (cosθ ≈ −0.52), implying a strong ordering at the interface as suggested by previous studies.19, 33, 60, 74, 8486 Note that since the principal eaks at 2875 cm−1 and 2938 cm−1 of the ssp SFG spectra in the C-H stretch region are ascribed to the methyl groups of the leucine residue (symmetric stretching and FR, respectively), the peaked distribution of Cγ-H supports the presence of an SFG-active ordered interface. The θ distributions for (D-) LK7β are essentially the same compared to the results for (L-) LK7β (Figure S9).

The distributions of the tilt angles over all the lysine residues in (L-) LK7β are presented in Figure 7b. The lysine sidechains preferentially face the water phase, as can be appreciated from the peaked Cα-Cβ distribution (blue) at θ = 180°. The distribution of Cα-H angles (black) is centered at θ = 90° consistent with the formation of anti-parallel β-sheets. The CβH2 moiety (green) is ordered. Note that the distribution of the CγH2 moieties (red) is bimodal, with peaks at θ ≈ 120° (cosθ ≈ −0.47) and θ ≈ 65° (cosθ ≈ 0.40). This implies that the alkyl chains are not fully ordered in perfect trans conformation and present some gauche defects, consistent with their VSFG optical activity.35 Again, the θ distributions for (D-) LK7β are unchanged (Figures S10 and S11).

The results for the tilt angle for one strand of (L-Leu)/(D-Lys) LK7β peptide are presented in Figures 7c and 7d. The leucine and lysine sidechains still point up and down, respectively, maintaining the tendency of hydrophobic or hydrophilic residues to point into the air or water phase (blue curves). The biggest contrasts with (L-) LK7β are the distributions of lysine CγH2 and Cα-H (Figure 7d, red and black curves). Both distributions become nearly isotropic, indicating that these modes should not be observed in a VSFG experiment.

Discussion

To the best of our knowledge, there have been no reports of VSFG being applied to study the effects of mixed (L-)/(D-) chirality on protein secondary structure. Here, we used achiral and chiral VSFG in combination with computational studies of protein dynamics to investigate biophysical properties of (L-Leu)/(D-Lys) LK7β at interfaces. We now provide an explanation for the achiral and chiral VSFG optical responses of (L-), (D-), and (L-Leu)/(D-Lys) LK7β variants, and relate chiral inversions of the peptide backbone with protein secondary structure and sidechain dynamics.

Identical Achiral VSFG Spectra of (L-) and (D-) LK7β.

The achiral VSFG spectra of (L-) and (D-) LK7β show strong resonant responses in the C-H stretching region around 2875 cm−1, 2913 cm−1, and 2938 cm−1 (Figures 1a and 2a). Our harmonic DFT frequency calculations of a model peptide, supplemented with anharmonic calculations of single amino acids, suggest assigning these peaks to leucine CH3 symmetric stretch, Fermi resonance modes arising from CH2 symmetric stretching (with possible contributions from lysine Cα-H stretching and CH3 symmetric stretch of the N-terminal acetyl cap), and leucine symmetric CH3 Fermi resonance, respectively, in agreement with previous studies.22, 33, 60, 7375 The spectral response of (L-) and (D-) LK7β in the C-H stretching region are identical, suggesting that a total inversion of the chirality of the amino acids does not affect the properties of the LK7β at the interface. These results are confirmed by molecular dynamics simulations demonstrating that both the distribution of backbone dihedral angles (Figure 6) as well as the orientations of sidechains (Figures S7 and S8) are identical for (L-) and (D-) LK7β peptide at the air-water interface.

A Silent CH2 Fermi Resonance Vibration in VSFG of (L-Leu)/(D-Lys) LK7β.

When the LK7β peptide is formed by an alternation of (L-Leu) and (D-Lys) amino acid residues, the vibrational mode at 2913 cm−1 in the achiral VSFG spectrum is silenced, although the vibrational modes at 2875 cm−1 and 2938 cm−1 still show (Figures 1b and 2b). This is especially intriguing given that linear, non-surface-selective spectroscopy such as ATR-FTIR (Figure S1) hardly distinguishes between the (L-), (D-), and (L-Leu)/(D-Lys) LK7β variants in the C-H stretching region.

In general, three possible conditions can lead to a silent achiral VSFG signal (Figure 8). (i) The molecular system adopts an ordered orientation that is VSFG inactive due to the polarization combination used in the experiment and, consequently, the selection rules over χ(2). For example, the dipole transition moment might be perpendicular to the applied electric field of the linearly-polarized IR beam. (ii) The molecular system adopts an ordered centrosymmetric orientation around the interface resulting in a (destructive) interference between vibrational modes and a null mean second-order polarizability,P¯(2)=0. (iii) The molecular system adopts a disordered, isotropic orientation around the interface, resulting in P¯(2)=0. In order to understand why the lysine CH2 2913 cm−1 peak is absent in the achiral VSFG spectrum of (L-Leu)/(D-Lys) LK7β, it is necessary to determine which of the three conditions is responsible for the silent VSFG signal. This required a combination of experimental and computational research, including DFT calculations of VSFG spectra and MD simulations.

Figure 8.

Figure 8.

Conditions for an inactive VSFG signal. For the purpose of example in the cases outlined, it is assumed that the permanent dipole moment points parallel to the bond.

Scenario (i): Reorientation.

Our Ramachandran analyses (Figure 6c and S6) show that (L-Leu)/(D-Lys) LK7β does not form stable β-sheets, but dynamically samples different conformations. On the other hand, the analysis of the orientation of CH2 groups of the lysine sidechain (Figure 7d) shows no apparent net orientation for this group. These results allow us to rule out scenario (i), namely that (L-Leu)/(D-Lys) LK7β at the interface adopts an ordered conformation that is differently oriented and thus VSFG inactive.

Scenario (ii): Centrosymmetric Order.

A similar analysis of our simulations also let us exclude scenario (ii), a centrosymmetric ordering at the interface. The Ramachandran free energy plot of (L-Leu)/(D-Lys) LK7β certainly demonstrates that the peptide backbone lacks secondary structure at the interface (Figure 6c). Moreover, the analysis of the orientation of lysine sidechains does not show an alternate orientation of CH2 groups that might lead to a destructive interference of vibrational modes. However, achiral VSFG reports not only on the absence or presence of protein secondary structure, but is convoluted with the vibrational modes of local chemical groups. This means that the free energy landscape of the peptide backbone structure is not enough to rule out all kinds of conformational order at the interface that could contribute (or silence) the VSFG response. Therefore, when analyzing achiral VSFG data it should be necessary to understand local bond tilt angle distributions as we have done in this work (Figure 7).

Scenario (iii): Isotropic Dynamics.

Our bond tilt angle analyses show that the introduction of chiral inversions at the lysine positions result in lysine Cα-H and CγH2 being isotropic (Figure 7d), while the orientation of leucine sidechains is not much affected by chiral inversion (Figure 7c). From this, we conclude that it is the isotropic nature of the coupled lysine Cα-H and CH2 groups in (L-Leu)/(D-Lys) LK7β that silences the response at 2913 cm−1, compared to the achiral VSFG spectra of (L-) LK7β and (D-) LK7β at interfaces.

Sidechain Preferences Dominate Protein Secondary Structure at Interfaces.

The structural free energy landscape of the peptide backbone of (L-Leu)/(D-Lys) LK7β shows multiple minima (Figure 6c) that are dynamically sampled by the peptide (Figure S6). This behavior might simply be attributed to the lack of inter-strand hydrogen bonding. However, our simulations show that even a single strand of (L-) LK7β adopts a conformation that resembles an extended β-strand structure (Figure 6d). Remarkably, the isolated strand of (L-) LK7β adopts a β-strand-like conformation even in the total absence of stabilizing inter-strand hydrogen bonds.

For all the LK7β variants studied here, tilt angle analyses show that the leucine sidechains point towards the air and the lysine sidechains point towards the water phase. This strong preference persists strongly even for the mixed chirality (L-Leu)/(D-Lys) system (Figure 7c and 7d). In line with previous studies,86 our results demonstrate that peptides adapt their architectures in line with the preferences of hydrophobic and hydrophilic sidechains towards the most stable phase even if these preferences incur large energetic costs by abolishing inter-strand hydrogen bonds. Our results go further, though: they also show that sidechain dynamics at the interface can be independent of chiral inversions along the peptide backbone.

Thus, our results suggest that interfacial (L-) and (D-) β-strand formation (and possibly β-sheet formation) can be driven from the alternating hydrophobicity and hydrophilicity of the peptide sidechains, more so than the disposition of the peptide backbone to form inter-strand hydrogen bonds. It is notable, however, that even this extremely strong alternating hydrophobic/hydrophilic preference can fail to form secondary structure if there is (L-)/(D-) heterochirality. Note that due to the chiral inversion in the lysine residues, the formation of an antiparallel β-sheet secondary structure would require the positioning of both leucine and lysine sidechains on the same side of the sheet. This arrangement of sidechains not only imposes steric hindrances on the β-sheet, due to the presence of bulky neighbor groups, but also prevents the hydrophobic or hydrophilic groups to be directed towards the more stable phase, completely destabilizing the β-sheet secondary structure at the interface. Because of the ability of VSFG to probe both local chemical functionalities and macroscopic chirality, these results can guide researcher expectations about the VSFG optical responses of proteins at interfaces.

Experimental Support that Chiral VSFG Response is Due to Macroscopic Chirality of β-sheets.

The achiral and chiral VSFG results in Figures 3a,b show that acid denaturation of (L-) LK7β and consequent loss of secondary structure does not silence the CH2 FR mode at 2913 cm−1. In fact, the achiral C-H stretch spectrum of (L-) LK7β is almost indistinguishable at pH = 7 or pH = 2 (Figure 3a). By contrast, the chiral amide I/amide II signal, which is indicative of anti-parallel β-sheets in (L-) and (D-) LK7β, is very strong at pH = 7 and almost completely abolished at pH = 2 (Figure 3b).

Intriguingly, our MD simulations of a single strand of (L-) LK7β (Figure 6d) offer insight about why the VSFG optical activities of the achiral C-H modes are insensitive to acid denaturation. The orientations of the C-H stretching modes are virtually unchanged when (L-) LK7β exists as an isolated β-strand (Figures S9 and S10) versus in anti-parallel β-sheets (Figures 7a and 7b). Based on this result and our vibrational mode assignments for the C-H stretching region, one could have expected that the achiral C-H spectrum at denatured pH = 2 versus native pH = 7 would be relatively indistinguishable.

In contrast, our research group previously showed that (L-) LK7β at the air-water interface shows no chiral C-H response at pH = 2.62 However, our MD simulations (Figure 6d) show that even one (L-) LK7β will form a β-strand due to the preference of its hydrophilic and hydrophobic sidechains to structure at the interface. Therefore, our studies lend critical support for the theory of Simpson4142 that suggests the chiral VSFG optical response of LK7β is due to macroscopic chirality of anti-parallel β-sheets.

Conclusions

We have applied VSFG spectroscopy in combination with DFT calculations and MD simulations to investigate the effects of chiral inversions of amino acids along the full length of a model anti-parallel β-sheet peptide at interfaces. We learned how chiral inversions perturb local structure of the peptide at the interface, and how such perturbations consequently disrupt the peptide secondary structure. The combination of VSFG experiments, DFT calculations of the VSFG optical response, and MD analyses of the peptide at the air-water interface show that, regardless of amino acid chirality, the peptide sidechains exhibit strong preferences for either the hydrophobic or hydrophilic phases. The peptide backbone adopts a conformation that accommodates these preferences within steric constraints. In the case of alternating (L-)/(D-) chirality in LK7β, for instance, this results in complete loss of the anti-parallel β-sheet secondary structure. Future studies that systematically address the effects of differently patterned (non-alternating) chiral inversions on the protein backbone, or extend these investigations to the study of protein-protein interfaces, will help to define the rules that link protein primary structure to secondary and tertiary structures. Such work should be particularly relevant for de novo design of unnatural protein structures with novel functions. Finally, controllable perturbations to protein molecular structure (e.g., chiral inversions, chemical and thermal denaturation, sidechain modification, etc.) can be employed as critical tests of chiral VSFG theory.

Supplementary Material

SI

Acknowledgments

E.A.P. is supported by the NIH (5T32GM008283-30). V.S.B. acknowledges support from the NIH Grant GM106121 and supercomputer time from the National Energy Research Scientific Computing Center (NERSC).

Footnotes

Supporting Information

ATR-FTIR of LK7β variants; anharmonic analyses of leucine and lysine residues; normal mode analyses of LK3β model; per residue free energy profiles of LK7β variants; time evolution of sidechain orientations of (L-) LK7β; strand dynamics of (L-Leu)/(D-Lys) LK7β; supplementary analyses of sidechain orientations.

References

  • 1.Ketchem RR; Hu W; Cross TA, High-resolution conformation of gramicidin A in a lipid bilayer by solid-state NMR. Science 1993, 261, 1457. [DOI] [PubMed] [Google Scholar]
  • 2.Nicholson LK; Cross TA, Gramicidin cation channel: an experimental determination of the right-handed helix sense and verification of .beta.-type hydrogen bonding. Biochemistry 1989, 28, 9379–9385. [DOI] [PubMed] [Google Scholar]
  • 3.Finking R; Marahiel MA, Biosynthesis of Nonribosomal Peptides. Annu. Rev. Microbiol. 2004, 58, 453–488. [DOI] [PubMed] [Google Scholar]
  • 4.Kessler N; Schuhmann H; Morneweg S; Linne U; Marahiel MA, The Linear Pentadecapeptide Gramicidin Is Assembled by Four Multimodular Nonribosomal Peptide Synthetases That Comprise 16 Modules with 56 Catalytic Domains. J. Biol. Chem. 2004, 279, 7413–7419. [DOI] [PubMed] [Google Scholar]
  • 5.Durani S, Protein Design with l- and d-α-Amino Acid Structures as the Alphabet. Acc. Chem. Res. 2008, 41, 1301–1308. [DOI] [PubMed] [Google Scholar]
  • 6.Huang P-S; Boyken SE; Baker D, The coming of age of de novo protein design. Nature 2016, 537, 320. [DOI] [PubMed] [Google Scholar]
  • 7.Lam KS; Salmon SE; Hersh EM; Hruby VJ; Kazmierski WM; Knapp RJ, A new type of synthetic peptide library for identifying ligand-binding activity. Nature 1991, 354, 82. [DOI] [PubMed] [Google Scholar]
  • 8.Oliva R; Chino M; Pane K; Pistorio V; De Santis A; Pizzo E; D’Errico G; Pavone V; Lombardi A; Del Vecchio P, et al. , Exploring the role of unnatural amino acids in antimicrobial peptides. Sci. Rep. 2018, 8, 8888. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Milton RC; Milton SC; Kent SB, Total chemical synthesis of a D-enzyme: the enantiomers of HIV-1 protease show reciprocal chiral substrate specificity [corrected]. Science 1992, 256, 1445. [DOI] [PubMed] [Google Scholar]
  • 10.Mahalakshmi R; Balaram P, In D-Amino Acids: A New Frontier in Amino Acid and Protein Research;, Konno R, Ed.; Nova Science Publisher: New York, 2007. [Google Scholar]
  • 11.Buri MV; Domingues TM; Paredes-Gamero EJ; Casaes-Rodrigues RL; Rodrigues EG; Miranda A, Resistance to Degradation and Cellular Distribution are Important Features for the Antitumor Activity of Gomesin. PLOS ONE 2013, 8, e80924. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Manabe T; Kawasaki K, D-form KLKLLLLLKLK-NH2 peptide exerts higher antimicrobial properties than its L-form counterpart via an association with bacterial cell wall components. Sci. Rep. 2017, 7, 43384. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Zerze GH; Khan MN; Stillinger FH; Debenedetti PG, Computational Investigation of the Effect of Backbone Chiral Inversions on Polypeptide Structure. J. Phys. Chem. B 2018, 122, 6357–6363. [DOI] [PubMed] [Google Scholar]
  • 14.Shen YR, Surface properties probed by second-harmonic and sum-frequency generation. Nature 1989, 337, 519. [Google Scholar]
  • 15.Shen L; Ulrich NW; Mello CM; Chen Z, Determination of conformation and orientation of immobilized peptides and proteins at buried interfaces. Chem. Phys. Lett. 2015, 619, 247–255. [Google Scholar]
  • 16.Wang Z; Morales-Acosta MD; Li S; Liu W; Kanai T; Liu Y; Chen Y-N; Walker FJ; Ahn CH; Leblanc RM, et al. , A narrow amide I vibrational band observed by sum frequency generation spectroscopy reveals highly ordered structures of a biofilm protein at the air/water interface. Chem. Commun. 2016, 52, 2956–2959. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Xiao D; Fu L; Liu J; Batista VS; Yan ECY, Amphiphilic Adsorption of Human Islet Amyloid Polypeptide Aggregates to Lipid/Aqueous Interfaces. J. Mol. Biol. 2012, 421, 537–547. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Yan ECY; Fu L; Wang Z; Liu W, Biological Macromolecules at Interfaces Probed by Chiral Vibrational Sum Frequency Generation Spectroscopy. Chem. Rev. 2014, 114, 8471–8498. [DOI] [PubMed] [Google Scholar]
  • 19.Fu L; Liu J; Yan ECY, Chiral Sum Frequency Generation Spectroscopy for Characterizing Protein Secondary Structures at Interfaces. J. Am. Chem. Soc. 2011, 133, 8094–8097. [DOI] [PubMed] [Google Scholar]
  • 20.Fu L; Wang Z; Psciuk BT; Xiao D; Batista VS; Yan ECY, Characterization of Parallel β-Sheets at Interfaces by Chiral Sum Frequency Generation Spectroscopy. J. Phys. Chem. Lett. 2015, 6, 1310–1315. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Fu L; Wang Z; Yan ECY, Chiral vibrational structures of proteins at interfaces probed by sum frequency generation spectroscopy. Int. J. Mol. Sci. 2011, 12, 9404–9425. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Weidner T; Apte JS; Gamble LJ; Castner DG, Probing the Orientation and Conformation of α-Helix and β-Strand Model Peptides on Self-Assembled Monolayers Using Sum Frequency Generation and NEXAFS Spectroscopy. Langmuir 2010, 26, 3433–3440. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Wang J; Chen X; Clarke ML; Chen Z, Detection of chiral sum frequency generation vibrational spectra of proteins and peptides at interfaces in situ. Proc. Natl. Acad. Sci. U. S. A. 2005, 102, 4978. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Shen YR, The Principles of Nonlinear Optics. Wiley-Interscience: New York, 1984. [Google Scholar]
  • 25.Boyd R, Nonlinear Optics. Academic Press: Cambridge, MA, 2008. [Google Scholar]
  • 26.Eisenthal KB, Liquid Interfaces Probed by Second-Harmonic and Sum-Frequency Spectroscopy. Chem. Rev. 1996, 96, 1343–1360. [DOI] [PubMed] [Google Scholar]
  • 27.Wang H-F; Gan W; Lu R; Rao Y; Wu B-H, Quantitative spectral and orientational analysis in surface sum frequency generation vibrational spectroscopy (SFG-VS). Int. Rev. Phys. Chem. 2005, 24, 191–256. [Google Scholar]
  • 28.Lambert AG; Davies PB; Neivandt DJ, Implementing the Theory of Sum Frequency Generation Vibrational Spectroscopy: A Tutorial Review. Appl. Spectrosc. Rev. 2005, 40, 103–145. [Google Scholar]
  • 29.Zhuang X; Miranda PB; Kim D; Shen YR, Mapping molecular orientation and conformation at interfaces by surface nonlinear optics. Phys. Rev. B 1999, 59, 12632–12640. [Google Scholar]
  • 30.Wang H-F; Velarde L; Gan W; Fu L, Quantitative Sum-Frequency Generation Vibrational Spectroscopy of Molecular Surfaces and Interfaces: Lineshape, Polarization, and Orientation. Annu. Rev. Phys. Chem. 2015, 66, 189–216. [DOI] [PubMed] [Google Scholar]
  • 31.Chen X; Sagle LB; Cremer PS, Urea Orientation at Protein Surfaces. J. Am. Chem. Soc. 2007, 129, 15104–15105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Schmüser L; Roeters S; Lutz H; Woutersen S; Bonn M; Weidner T, Determination of Absolute Orientation of Protein α-Helices at Interfaces Using Phase-Resolved Sum Frequency Generation Spectroscopy. J. Phys. Chem. Lett. 2017, 8, 3101–3105. [DOI] [PubMed] [Google Scholar]
  • 33.Ji N; Shen Y-R, Sum frequency vibrational spectroscopy of leucine molecules adsorbed at air–water interface. J. Chem. Phys. 2004, 120, 7107–7112. [DOI] [PubMed] [Google Scholar]
  • 34.Ma G; Allen HC, DPPC Langmuir Monolayer at the Air−Water Interface:  Probing the Tail and Head Groups by Vibrational Sum Frequency Generation Spectroscopy. Langmuir 2006, 22, 5341–5349. [DOI] [PubMed] [Google Scholar]
  • 35.Guyot-Sionnest P; Hunt JH; Shen YR, Sum-frequency vibrational spectroscopy of a Langmuir film: Study of molecular orientation of a two-dimensional system. Phys. Rev. Lett. 1987, 59, 1597–1600. [DOI] [PubMed] [Google Scholar]
  • 36.Himmelhaus M; Eisert F; Buck M; Grunze M, Self-Assembly of n-Alkanethiol Monolayers. A Study by IR−Visible Sum Frequency Spectroscopy (SFG). J. Phys. Chem. B 2000, 104, 576–584. [Google Scholar]
  • 37.Lu R; Gan W; Wu B.-h.; Chen H; Wang H.-f., Vibrational Polarization Spectroscopy of CH Stretching Modes of the Methylene Group at the Vapor/Liquid Interfaces with Sum Frequency Generation. J. Phys. Chem. B 2004, 108, 7297–7306. [Google Scholar]
  • 38.Lu R; Gan W; Wu B.-h.; Zhang Z; Guo Y; Wang H.-f., C−H Stretching Vibrations of Methyl, Methylene and Methine Groups at the Vapor/Alcohol (n = 1−8) Interfaces. J. Phys. Chem. B 2005, 109, 14118–14129. [DOI] [PubMed] [Google Scholar]
  • 39.Fu L; Zhang Y; Wei Z-H; Wang H-F, Intrinsic Chirality and Prochirality at Air/R-(+)- and S-(−)-Limonene Interfaces: Spectral Signatures With Interference Chiral Sum-Frequency Generation Vibrational Spectroscopy. Chirality 2014, 26, 509–520. [DOI] [PubMed] [Google Scholar]
  • 40.Wang H-F, In Comprehensive Chiroptical Spectroscopy; Berova N, Polavarapu PL, Nakanishi K Woody RW, Eds.; John Wiley & Sons, Inc.: Hoboken, 2011. [Google Scholar]
  • 41.Moad AJ; Simpson GJ, A Unified Treatment of Selection Rules and Symmetry Relations for Sum-Frequency and Second Harmonic Spectroscopies. J. Phys. Chem. B 2004, 108, 3548–3562. [Google Scholar]
  • 42.Simpson GJ, Molecular Origins of the Remarkable Chiral Sensitivity of Second-Order Nonlinear Optics. ChemPhysChem 2004, 5, 1301–1310. [DOI] [PubMed] [Google Scholar]
  • 43.Ma G; Liu J; Fu L; Yan ECY, Probing Water and Biomolecules at the Air—Water Interface with a Broad Bandwidth Vibrational Sum Frequency Generation Spectrometer from 3800 to 900 cm−1. Appl. Spectrosc. 2009, 63, 528–537. [DOI] [PubMed] [Google Scholar]
  • 44.Best RB; Zhu X; Shim J; Lopes PEM; Mittal J; Feig M; MacKerell AD, Optimization of the Additive CHARMM All-Atom Protein Force Field Targeting Improved Sampling of the Backbone ϕ, ψ and Side-Chain χ1 and χ2 Dihedral Angles. J. Chem. Theory Comput. 2012, 8, 3257–3273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Jorgensen WL; Chandrasekhar J; Madura JD; Impey RW; Klein ML, Comparison of simple potential functions for simulating liquid water. J. Chem. Phys. 1983, 79, 926–935. [Google Scholar]
  • 46.Phillips JC; Braun R; Wang W; Gumbart J; Tajkhorshid E; Villa E; Chipot C; Skeel RD; Kalé L; Schulten K, Scalable molecular dynamics with NAMD. J. Comput. Chem. 2005, 26, 1781–1802. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Miyamoto S; Kollman PA, Settle: An analytical version of the SHAKE and RATTLE algorithm for rigid water models. J. Comput. Chem. 1992, 13, 952–962. [Google Scholar]
  • 48.Humphrey W; Dalke A; Schulten K, VMD: Visual molecular dynamics. J. Mol. Graphics 1996, 14, 33–38. [DOI] [PubMed] [Google Scholar]
  • 49.Frisch MJ; Trucks GW; Schlegel HB; Scuseria GE; Robb MA; Cheeseman JR; Scalmani G; Barone V; Petersson GA; Nakatsuji H, et al. Gaussian 09 Rev. E.01, Wallingford, CT, 2009. [Google Scholar]
  • 50.Becke AD, Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar]
  • 51.Hariharan PC; Pople JA, The influence of polarization functions on molecular orbital hydrogenation energies. Theor. Chim. Acta 1973, 28, 213–222. [Google Scholar]
  • 52.Anfuso CL; Snoeberger RC; Ricks AM; Liu W; Xiao D; Batista VS; Lian T, Covalent Attachment of a Rhenium Bipyridyl CO2 Reduction Catalyst to Rutile TiO2. J. Am. Chem. Soc. 2011, 133, 6922–6925. [DOI] [PubMed] [Google Scholar]
  • 53.Anfuso CL; Xiao D; Ricks AM; Negre CFA; Batista VS; Lian T, Orientation of a Series of CO2 Reduction Catalysts on Single Crystal TiO2 Probed by Phase-Sensitive Vibrational Sum Frequency Generation Spectroscopy (PS-VSFG). J. Phys. Chem. C 2012, 116, 24107–24114. [Google Scholar]
  • 54.Fu L; Xiao D; Wang Z; Batista VS; Yan ECY, Chiral Sum Frequency Generation for In Situ Probing Proton Exchange in Antiparallel β-Sheets at Interfaces. J. Am. Chem. Soc. 2013, 135, 3592–3598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Clark ML; Rudshteyn B; Ge A; Chabolla SA; Machan CW; Psciuk BT; Song J; Canzi G; Lian T; Batista VS, et al. , Orientation of Cyano-Substituted Bipyridine Re(I) fac-Tricarbonyl Electrocatalysts Bound to Conducting Au Surfaces. J. Phys. Chem. C 2016, 120, 1657–1665. [Google Scholar]
  • 56.Ge A; Rudshteyn B; Psciuk BT; Xiao D; Song J; Anfuso CL; Ricks AM; Batista VS; Lian T, Surface-Induced Anisotropic Binding of a Rhenium CO2-Reduction Catalyst on Rutile TiO2(110) Surfaces. J. Phys. Chem. C 2016, 120, 20970–20977. [Google Scholar]
  • 57.Ge A; Videla PE; Lee GL; Rudshteyn B; Song J; Kubiak CP; Batista VS; Lian T, Interfacial Structure and Electric Field Probed by in Situ Electrochemical Vibrational Stark Effect Spectroscopy and Computational Modeling. J. Phys. Chem. C 2017, 121, 18674–18682. [Google Scholar]
  • 58.Ge A; Videla PE; Rudshteyn B; Liu Q; Batista VS; Lian T, Dopant-Dependent SFG Response of Rhenium CO2 Reduction Catalysts Chemisorbed on SrTiO3 (100) Single Crystals. J. Phys. Chem. C 2018, 122, 13944–13952. [Google Scholar]
  • 59.DeGrado WF; Lear JD, Induction of peptide conformation at apolar water interfaces. 1. A study with model peptides of defined hydrophobic periodicity. J. Am. Chem. Soc. 1985, 107, 7684–7689. [Google Scholar]
  • 60.Phillips DC; York RL; Mermut O; McCrea KR; Ward RS; Somorjai GA, Side Chain, Chain Length, and Sequence Effects on Amphiphilic Peptide Adsorption at Hydrophobic and Hydrophilic Surfaces Studied by Sum-Frequency Generation Vibrational Spectroscopy and Quartz Crystal Microbalance. J. Phys. Chem. C 2007, 111, 255–261. [Google Scholar]
  • 61.Stokes GY; Gibbs-Davis JM; Boman FC; Stepp BR; Condie AG; Nguyen ST; Geiger FM, Making “Sense” of DNA. J. Am. Chem. Soc. 2007, 129, 7492–7493. [DOI] [PubMed] [Google Scholar]
  • 62.Wang Z; Fu L; Yan ECY, C–H Stretch for Probing Kinetics of Self-Assembly into Macromolecular Chiral Structures at Interfaces by Chiral Sum Frequency Generation Spectroscopy. Langmuir 2013, 29, 4077–4083. [DOI] [PubMed] [Google Scholar]
  • 63.Bloino J; Barone V, A second-order perturbation theory route to vibrational averages and transition properties of molecules: General formulation and application to infrared and vibrational circular dichroism spectroscopies. J. Chem. Phys. 2012, 136, 124108. [DOI] [PubMed] [Google Scholar]
  • 64.Yagi K; Hirata S; Hirao K, Vibrational quasi-degenerate perturbation theory: applications to fermi resonance in CO2, H2CO, and C6H6. Phys. Chem. Chem. Phys. 2008, 10, 1781–1788. [DOI] [PubMed] [Google Scholar]
  • 65.Barone V; Bloino J; Guido CA; Lipparini F, A fully automated implementation of VPT2 Infrared intensities. Chem. Phys. Lett. 2010, 496, 157–161. [Google Scholar]
  • 66.Biczysko M; Bloino J; Carnimeo I; Panek P; Barone V, Fully ab initio IR spectra for complex molecular systems from perturbative vibrational approaches: Glycine as a test case. J. Mol. Struct. 2012, 1009, 74–82. [Google Scholar]
  • 67.Bowman JM; Carrington T; Meyer H-D, Variational quantum approaches for computing vibrational energies of polyatomic molecules. Mol. Phys. 2008, 106, 2145–2182. [Google Scholar]
  • 68.Roy TK; Gerber RB, Vibrational self-consistent field calculations for spectroscopy of biological molecules: new algorithmic developments and applications. Phys. Chem. Chem. Phys. 2013, 15, 9468–9492. [DOI] [PubMed] [Google Scholar]
  • 69.Qu C; Bowman JM, Quantum approaches to vibrational dynamics and spectroscopy: is ease of interpretation sacrificed as rigor increases? Phys. Chem. Chem. Phys. 2018. [DOI] [PubMed] [Google Scholar]
  • 70.Ho J; Psciuk BT; Chase HM; Rudshteyn B; Upshur MA; Fu L; Thomson RJ; Wang H-F; Geiger FM; Batista VS, Sum Frequency Generation Spectroscopy and Molecular Dynamics Simulations Reveal a Rotationally Fluid Adsorption State of α-Pinene on Silica. J. Phys. Chem. C 2016, 120, 12578–12589. [Google Scholar]
  • 71.Jetzki M; Luckhaus D; Signorell R, Fermi resonance and conformation in glycolaldehyde particles. Can. J. Chem. 2004, 82, 915–924. [Google Scholar]
  • 72.Guo Z; Zheng W; Hamoudi H; Dablemont C; Esaulov VA; Bourguignon B, On the chain length dependence of CH3 vibrational mode relative intensities in sum frequency generation spectra of self assembled alkanethiols. Surf. Sci. 2008, 602, 3551–3559. [Google Scholar]
  • 73.York RL; Browne WK; Geissler PL; Somorjai GA, Peptides Adsorbed on Hydrophobic Surfaces—A Sum Frequency Generation Vibrational Spectroscopy and Modeling Study. Isr. J. Chem. 2010, 47, 51–58. [Google Scholar]
  • 74.Watry MR; Richmond GL, Orientation and Conformation of Amino Acids in Monolayers Adsorbed at an Oil/Water Interface As Determined by Vibrational Sum-Frequency Spectroscopy. J. Phys. Chem. B 2002, 106, 12517–12523. [Google Scholar]
  • 75.Holinga GJ; York RL; Onorato RM; Thompson CM; Webb NE; Yoon AP; Somorjai GA, An SFG Study of Interfacial Amino Acids at the Hydrophilic SiO2 and Hydrophobic Deuterated Polystyrene Surfaces. J. Am. Chem. Soc. 2011, 133, 6243–6253. [DOI] [PubMed] [Google Scholar]
  • 76.York RL; Mermut O; Phillips DC; McCrea KR; Ward RS; Somorjai GA, Influence of Ionic Strength on the Adsorption of a Model Peptide on Hydrophilic Silica and Hydrophobic Polystyrene Surfaces:  Insight from SFG Vibrational Spectroscopy. J. Phys. Chem. C 2007, 111, 8866–8871. [Google Scholar]
  • 77.Mermut O; Phillips DC; York RL; McCrea KR; Ward RS; Somorjai GA, In Situ Adsorption Studies of a 14-Amino Acid Leucine-Lysine Peptide onto Hydrophobic Polystyrene and Hydrophilic Silica Surfaces Using Quartz Crystal Microbalance, Atomic Force Microscopy, and Sum Frequency Generation Vibrational Spectroscopy. J. Am. Chem. Soc. 2006, 128, 3598–3607. [DOI] [PubMed] [Google Scholar]
  • 78.Towse C-L; Hopping G; Vulovic I; Daggett V, Nature versus design: the conformational propensities of d-amino acids and the importance of side chain chirality. Prot. Eng., Des. Sel. 2014, 27, 447–455. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Fairman R; Anthony-Cahill SJ; DeGrado WF, The helix-forming propensity of D-alanine in a right-handed .alpha.-helix. J. Am. Chem. Soc. 1992, 114, 5458–5459. [Google Scholar]
  • 80.Hermans J; Anderson AG; Yun RH, Differential helix propensity of small apolar side chains studied by molecular dynamics simulations. Biochemistry 1992, 31, 5646–5653. [DOI] [PubMed] [Google Scholar]
  • 81.Krause E; Beyermann M; Dathe M; Rothemund S; Bienert M, Location of an Amphipathic .alpha.-Helix in Peptides Using Reversed-Phase HPLC Retention Behavior of D-Amino Acid Analogs. Anal. Chem. 1995, 67, 252–258. [DOI] [PubMed] [Google Scholar]
  • 82.Krause E; Bienert M; Schmieder P; Wenschuh H, The Helix-Destabilizing Propensity Scale of d-Amino Acids:  The Influence of Side Chain Steric Effects. J. Am. Chem. Soc. 2000, 122, 4865–4870. [Google Scholar]
  • 83.Nanda V; DeGrado WF, Simulated Evolution of Emergent Chiral Structures in Polyalanine. J. Am. Chem. Soc. 2004, 126, 14459–14467. [DOI] [PubMed] [Google Scholar]
  • 84.Weidner T; Breen NF; Li K; Drobny GP; Castner DG, Sum frequency generation and solid-state NMR study of the structure, orientation, and dynamics of polystyrene-adsorbed peptides. Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 13288. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Donovan MA; Yimer YY; Pfaendtner J; Backus EHG; Bonn M; Weidner T, Ultrafast Reorientational Dynamics of Leucine at the Air–Water Interface. J. Am. Chem. Soc. 2016, 138, 5226–5229. [DOI] [PubMed] [Google Scholar]
  • 86.Donovan MA; Lutz H; Yimer YY; Pfaendtner J; Bonn M; Weidner T, LK peptide side chain dynamics at interfaces are independent of secondary structure. Phys. Chem. Chem. Phys. 2017, 19, 28507–28511. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

SI

RESOURCES