10
Exploiting ChromophoreProtein Interactions through Linker Engineering To Tune Photoinduced Dynamics in a Biomimetic Light- Harvesting Platform Milan Delor, ,Jing Dai, ,Trevor D. Roberts, ,Julia R. Rogers, ,Samia M. Hamed, Jerey B. Neaton, ,§,Phillip L. Geissler, ,Matthew B. Francis, ,# and Naomi S. Ginsberg* ,,,§,#,Department of Chemistry and Department of Physics, University of California Berkeley, Berkeley, California 94720, United States § Kavli Energy NanoSciences Institute, Berkeley, California 94720, United States Molecular Foundry, Chemical Sciences Division, # Material Sciences Division, and Molecular Biophysics and Integrated Bioimaging Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, United States * S Supporting Information ABSTRACT: Creating articial systems that mimic and surpass those found in nature is one of the great challenges of modern science. In the context of photosynthetic light harvesting, the diculty lies in attaining utmost control over the energetics, positions and relative orientations of chromophores in densely packed arrays to transfer electronic excitation energy to desired locations with high eciency. Toward achieving this goal, we use a highly versatile biomimetic protein scaold from the tobacco mosaic virus coat protein on which chromophores can be attached at precise locations via linkers of diering lengths and rigidities. We show that minor linker modications, including switching chiral congurations and alkyl chain shortening, lead to signicant lengthening of the ultrafast excited state dynamics of the system as the linkers are shortened and rigidied. Molecular dynamics simulations provide molecular-level detail over how the chromophore attachment orientations, positions, and distances from the protein surface lead to the observed trends in system dynamics. In particular, we nd that short and rigid linkers are able to sandwich water molecules between chromophore and protein, leading to chromophorewaterprotein supracomplexes with intricately coupled dynamics that are highly dependent on their local protein environment. In addition, cyclohexyl-based linkers are identied as ideal candidates to retain rotational correlations over several nanoseconds and thus lock relative chromophore orientations throughout the lifetime of an exciton. Combining linker engineering with judicious placement of chromophores on the hydrated protein scaold to exploit dierent chromophorebath couplings provides a clear and eective path to producing highly controllable articial light-harvesting systems that can increasingly mimic their natural counterparts, thus aiding to elucidate natural photosynthetic mechanisms. INTRODUCTION Electronic energy transfer to redox reaction centers in photosynthetic organisms can approach near-perfect quantum eciencies by virtue of highly evolved chromophoreprotein structures with delicately tuned and optimized spatioenergetic landscapes. 14 Much has yet to be learned by systematically elucidating the factors governing their exceptional properties; unfortunately, perturbing the intricate congurations of photo- synthetic organisms by removing chromophores or mutating protein residues often destabilizes the overall molecular architecture or alters multiple variables at a time. The resulting diculty in performing controlled experiments obfuscates how function emerges from these complex structures. One alternative strategy to learn about nature is to create biomimetic light-harvesting assemblies using modular scaolds that enable the systematic changing of one structural variable at a time. 515 In practice, however, the energy collection and transfer eciencies over nanometer-to-micron length-scales achieved in nature are extremely dicult to reproduce in articial systems. This challenge is largely due to the requirement that synthetic platforms both accommodate large chromophore densities and establish the precise positioning and energetic properties that protect fragile electronic excitations and optimize excitation energy transfer. 1621 Although much progress has been made on biomimetic light- harvesting scaolds, control of the specic conguration of chromophores relative to one another and relative to the protein itself has been lacking and has diered substantially from natural light-harvesting complexes, where bound Received: December 22, 2017 Published: May 9, 2018 Article pubs.acs.org/JACS Cite This: J. Am. Chem. Soc. 2018, 140, 6278-6287 © 2018 American Chemical Society 6278 DOI: 10.1021/jacs.7b13598 J. Am. Chem. Soc. 2018, 140, 62786287 Downloaded by UNIV OF CALIFORNIA BERKELEY at 09:02:03:469 on June 11, 2019 from https://pubs.acs.org/doi/10.1021/jacs.7b13598.

Exploiting Chromophore–Protein Interactions through Linker …geisslergroup.org/wp-content/uploads/2019/06/Pub2018-5.pdf · 2019-06-11 · shows the overlaid optimized ground and

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Exploiting Chromophore–Protein Interactions through Linker …geisslergroup.org/wp-content/uploads/2019/06/Pub2018-5.pdf · 2019-06-11 · shows the overlaid optimized ground and

Exploiting Chromophore−Protein Interactions through LinkerEngineering To Tune Photoinduced Dynamics in a Biomimetic Light-Harvesting PlatformMilan Delor,†,⊗ Jing Dai,†,⊗ Trevor D. Roberts,†,⊗ Julia R. Rogers,†,⊗ Samia M. Hamed,†

Jeffrey B. Neaton,‡,§,∥ Phillip L. Geissler,†,⊥ Matthew B. Francis,†,# and Naomi S. Ginsberg*,†,‡,§,#,∇

†Department of Chemistry and ‡Department of Physics, University of California Berkeley, Berkeley, California 94720, United States§Kavli Energy NanoSciences Institute, Berkeley, California 94720, United States∥Molecular Foundry, ⊥Chemical Sciences Division, #Material Sciences Division, and ∇Molecular Biophysics and IntegratedBioimaging Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, United States

*S Supporting Information

ABSTRACT: Creating artificial systems that mimic and surpassthose found in nature is one of the great challenges of modernscience. In the context of photosynthetic light harvesting, thedifficulty lies in attaining utmost control over the energetics,positions and relative orientations of chromophores in denselypacked arrays to transfer electronic excitation energy to desiredlocations with high efficiency. Toward achieving this goal, we usea highly versatile biomimetic protein scaffold from the tobaccomosaic virus coat protein on which chromophores can beattached at precise locations via linkers of differing lengths andrigidities. We show that minor linker modifications, includingswitching chiral configurations and alkyl chain shortening, lead tosignificant lengthening of the ultrafast excited state dynamics ofthe system as the linkers are shortened and rigidified. Moleculardynamics simulations provide molecular-level detail over how the chromophore attachment orientations, positions, and distancesfrom the protein surface lead to the observed trends in system dynamics. In particular, we find that short and rigid linkers are ableto sandwich water molecules between chromophore and protein, leading to chromophore−water−protein supracomplexes withintricately coupled dynamics that are highly dependent on their local protein environment. In addition, cyclohexyl-based linkersare identified as ideal candidates to retain rotational correlations over several nanoseconds and thus lock relative chromophoreorientations throughout the lifetime of an exciton. Combining linker engineering with judicious placement of chromophores onthe hydrated protein scaffold to exploit different chromophore−bath couplings provides a clear and effective path to producinghighly controllable artificial light-harvesting systems that can increasingly mimic their natural counterparts, thus aiding toelucidate natural photosynthetic mechanisms.

■ INTRODUCTION

Electronic energy transfer to redox reaction centers inphotosynthetic organisms can approach near-perfect quantumefficiencies by virtue of highly evolved chromophore−proteinstructures with delicately tuned and optimized spatioenergeticlandscapes.1−4 Much has yet to be learned by systematicallyelucidating the factors governing their exceptional properties;unfortunately, perturbing the intricate configurations of photo-synthetic organisms by removing chromophores or mutatingprotein residues often destabilizes the overall moleculararchitecture or alters multiple variables at a time. The resultingdifficulty in performing controlled experiments obfuscates howfunction emerges from these complex structures.One alternative strategy to learn about nature is to create

biomimetic light-harvesting assemblies using modular scaffoldsthat enable the systematic changing of one structural variable at

a time.5−15 In practice, however, the energy collection andtransfer efficiencies over nanometer-to-micron length-scalesachieved in nature are extremely difficult to reproduce inartificial systems. This challenge is largely due to therequirement that synthetic platforms both accommodate largechromophore densities and establish the precise positioningand energetic properties that protect fragile electronicexcitations and optimize excitation energy transfer.16−21

Although much progress has been made on biomimetic light-harvesting scaffolds, control of the specific configuration ofchromophores relative to one another and relative to theprotein itself has been lacking and has differed substantiallyfrom natural light-harvesting complexes, where bound

Received: December 22, 2017Published: May 9, 2018

Article

pubs.acs.org/JACSCite This: J. Am. Chem. Soc. 2018, 140, 6278−6287

© 2018 American Chemical Society 6278 DOI: 10.1021/jacs.7b13598J. Am. Chem. Soc. 2018, 140, 6278−6287

Dow

nloa

ded

by U

NIV

OF

CA

LIF

OR

NIA

BE

RK

EL

EY

at 0

9:02

:03:

469

on J

une

11, 2

019

from

http

s://p

ubs.

acs.

org/

doi/1

0.10

21/ja

cs.7

b135

98.

Page 2: Exploiting Chromophore–Protein Interactions through Linker …geisslergroup.org/wp-content/uploads/2019/06/Pub2018-5.pdf · 2019-06-11 · shows the overlaid optimized ground and

chromophores often fit tightly into protein pockets.22−25 Mostbiomimetic attempts employ dyes used in fluorescentbioimaging, where long and flexible tethers connecting thechromophore to the protein scaffold are common.10,26,27 Suchbioconjugation schemes often preclude precise control overpositioning and orientation, as the chromophores can samplelarge and random conformational volumes. A clear needtherefore exists for crafting chromophore−protein linkers thatcan establish and maintain interchromophore orientations andseparations through conformational restrictions.Here we address the need for well-defined chromophore

configurations by comparing the ultrafast transient absorption(TA) time scales and results from molecular dynamics (MD)simulations of chromophores bound to supramolecular proteinassemblies derived from the tobacco mosaic virus (TMV) coatprotein via a series of increasingly constricting linkers. Self-assembled structures based on TMV coat proteins have beenshown to be exceptionally versatile platforms for biomimeticlight-harvesting assemblies.9−12,28,29 In this work, we show thatsmall changes in linker length, rigidity, and chiral configurationcan significantly extend the excited state dynamics of thesystem. By finely tuning the position and range of motion ofchromophores, we demonstrate an unprecedented level ofcontrol over chromophore−protein−solvent interactions andidentify with great precision the mechanisms responsible for thelinker-mediated dynamic slow-down of excited state processes.We identify the most rigid linkers as ideal for biomimetic lightharvesting, as they both retard chromophore energy-lossprocesses and retain orientational correlations throughout theexcited state lifetime of the chromophore. Our results highlightthe crucial but typically overlooked importance of usingpurpose-specific linkers to emulate the dynamic tuning ofspatioenergetic landscapes in naturally occurring light-harvest-ing assemblies. They suggest a simple but powerful design

principle to balance chromophore−bath and interchromophorecouplings to achieve specific light-induced functions.

■ RESULTS AND DISCUSSION

To explore our ability to control chromophore−proteininteractions for biomimetic light-harvesting, we developed anew library of linkers of varying lengths and rigidities to tether achromophore to one of two specific positions on the TMVprotein with differing degrees of confinement. In this study, weuse the circular permutant of TMV (cpTMV), which self-assembles into an 18 nm-wide C2 symmetric double-ringstructure. By introducing a uniquely reactive cysteine residue,we are able to attach chromophores covalently at specificlocations on the protein monomers, which can then self-assemble in different supramolecular configurations dependingon the pH and ionic strength conditions. The assembly consistsof 34 monomers (17 per ring) with a hydrated cavity extendingradially outward from the central pore (Figure 1a,b).11,12 Wecombine this versatile platform with chromophore−proteinlinker engineering to conduct a systematic study of the effectsof conformational constriction and attachment orientation onchromophore photophysics. Because a major aim is to mimicenvironmentally protected photosynthetic chromophores, westudy systems where only one chromophore is present percpTMV assembly, thus allowing detailed characterization of thechromophore−protein interactions, free of complicating factorssuch as interchromophore interactions. We use ultrafast TA tocharacterize the excited state dynamics of each system, focusingon solvation and chromophore intramolecular structuralreorganization occurring within picoseconds of light absorption.MD simulations with explicit inclusion of the protein scaffoldand residues, the surrounding water, and the chromophore thenallow us to rationalize the trends in excited state dynamics

Figure 1. Biomimetic light-harvesting assembly comprised of the cpTMV scaffold labeled by a single chromophore via different linkers studied here.(a) A side-view cross-section of the cpTMV assembly, displaying C2 symmetry with a hydrated cavity at the center of the assembly. (b) A top view ofthe full assembly, with a single monomer highlighted as well as the mutated S23C residue (yellow) to which a chromophore can be conjugated. (c)Top-bottom monomer pairs, illustrating the S23C and Q101C labeling positions (yellow and pale blue, respectively), with a chromophoreconjugated for scale. (d) Sulforhodamine B (SRB) is linked to a maleimide for bioconjugation to cysteine residues via the different linkers shown in(e). The red, green, orange, and blue color scheme for linkers are maintained throughout the manuscript.

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.7b13598J. Am. Chem. Soc. 2018, 140, 6278−6287

6279

Page 3: Exploiting Chromophore–Protein Interactions through Linker …geisslergroup.org/wp-content/uploads/2019/06/Pub2018-5.pdf · 2019-06-11 · shows the overlaid optimized ground and

extracted from TA in terms of specific interactions betweenchromophore and protein bath, depending on the linker used.The attachment positions, chromophore and series of linkers

used to study these protein−chromophore interactions aredisplayed in Figure 1c−e. As shown in Figure 1c, two distinctenvironments are studied: the protein−water interface(chromophore labeling site S23C) and the nanoscale hydratedprotein cavity (site Q101C). The chromophore in question isSulforhodamine B (SRB, Figure 1d), chosen for its high visiblelight absorption cross-section (ε570 nm ≈ 83 000 M−1 cm−1 inwater) and relative rigidity due to few rotational degrees offreedom in the electronically active moieties (see below).Figure 1e shows the series of linkers used to control thedistance of the chromophores from the protein surface andtheir conformational volume. In all cases, these linkers areshorter and more rigid than in typical bioimaging labels, whichuse pentyl or longer alkyl chain linkers. In our series, changingfrom a 4-carbon butyl (Bu) to a 2-carbon ethyl (Et) moietystraightforwardly shortens the linker. Allylic 1,3-strain whenusing cyclohexyl (Cyc) moieties further reduces conformationalfreedom compared to alkyl chains. The chiral center of theselinkers provides an additional degree of control, as the twoenantiomers (Cyc-SS and Cyc-RR) are expected to lead to

different attachment orientations on the protein surface.Altogether 12 systems are investigated: SRB attached to twoseparate positions (S23C and Q101C, Figure 1c) on cpTMVvia four different linkers (Bu, Et, Cyc-SS, Cyc-RR, Figure 1e),along with control experiments using free maleimide-function-alized chromophores after reaction with the sodium salt of 2-mercaptoethanesulfonate in buffer. All experiments areperformed in solution using a sodium phosphate buffer.These 12 systems form a series with systematically variedchromophore−bath couplings, enabling extraction of detailedinformation about how the protein environment affects thenuclear-electronic dynamics of each assembly.Figure 2a displays the TA spectra of SRB-Cyc-SS attached at

the outside S23C site of cpTMV at representative delay timesfollowing light excitation at 570 nm, the lowest allowedelectronic transition. The spectra display excited stateabsorption (ESA) around 450 nm and a band convolvingground state bleaching (GSB) and stimulated emission (SE)centered at 580 nm. The TA spectra for all 12 systems followvery similar spectral profiles and are shown in the SupportingInformation, along with details of the global kinetic analysisused to extract the time scales associated with the excited stateevolution of each system (Figures S2−S4). In every case, the

Figure 2. Excited state dynamics of SRB following 570 nm, ∼40 fs laser light excitation. (a) TA spectra at representative delay times between 1 psand 1.2 ns for SRB-Cyc-SS attached at the S23C (outside) position of cpTMV. The normalized linear absorption and emission spectra are plottedfor comparison. (b) Decay associated spectra (DAS) for the two time components needed to fit the data in (a). The shorter, 9.9 ps, component(red) displays two dispersive differential profiles characteristic of (1) redshifting ESA (positive/negative) and (2) redshifting SE (negative/positive)around 450 and 600 nm, respectively assigned to intramolecular structural reorganization and solvation of the chromophore’s excited state. The insetshows the overlaid optimized ground and excited state structures as calculated from DFT, displaying minor intramolecular structural rearrangement.(c) TD-DFT calculations at the PBE0/6-31+g* level including solvation effects from the IEF-PCM have yielded vertical transition energies from theground state to the bright S1 state and dark S2 state as a function of phenyl-xanthene rotation angle. The region of the excited state surface that iseffectively inaccessible at room temperature (RT) is shaded light gray. Although an excited state crossing exists around a phenyl-xanthene rotationangle of 55 degrees, this configuration is inaccessible at RT. For comparison, the ground state energies are plotted in gray, with energies that rapidlyincrease past a rotation of 10 degrees from an orthogonal geometry. (d) Short time-component TA lifetimes corresponding to the red DAS in (b)for all chromophore-linker compounds in buffer and at their respective attachment sites on cpTMV.

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.7b13598J. Am. Chem. Soc. 2018, 140, 6278−6287

6280

Page 4: Exploiting Chromophore–Protein Interactions through Linker …geisslergroup.org/wp-content/uploads/2019/06/Pub2018-5.pdf · 2019-06-11 · shows the overlaid optimized ground and

kinetics across the visible spectrum are well fit by abiexponential function, with a short component on the orderof a few picoseconds and a longer component on the order of afew nanoseconds. Although the spectral evolution is equivalentacross systems, indicating the protein environment does notchange the nature of the populated excited states, the timescales for these processes are vastly different, providing a usefulhandle on the influence of the biomolecular environment onchromophore-bath dynamics.Before comparing results across the 12 different compounds,

we provide a brief interpretation of the observed chromophoredynamics themselves, using time-dependent density functionaltheory (TD-DFT) to support our excited state assignments.Although the excited state dynamics of rhodamine dyes arewell-studied, the excited state processes can depend on thespecific molecular details of both the phenyl and xanthenemoieties.30−35 Solvent and intramolecular structural reorganiza-tion are expected to occur within the first few picoseconds afterphotoexcitation. Intramolecular reorganization in rhodamines isnot yet fully characterized, but involves some combination ofthe following: rotation of the dialkylamino groups and/or C−Nbond order reduction, rotation of the initially orthogonalphenyl ring with respect to the xanthene core, and/or bucklingof the xanthene moiety.30,33,36 This reorganization can facilitateelectron delocalization in the excited state. Previous TD-DFTcalculations on Rhodamine B, a close analogue to SRB with acarboxylate moiety replacing the SO3

− group on the phenylring, identified a dark, low-lying charge transfer (CT) state thatmay be responsible for partial quenching of the lowest-lyingbright state.37 Our TD-DFT calculations corroborate theexistence of a lower-lying CT state when the phenyl-xanthenedihedral approaches a planar configuration in SRB. Accessingthis CT state would, however, necessitate rotation of the phenylring with respect to the xanthene core towards a quasi-planargeometry, representing an energetic barrier of ∼2 eV on theexcited state potential energy surface from the initiallyorthogonal geometry (Figure 2c). This barrier arises due tosevere steric constraints imposed by the SO3

− group, renderingthis quenching pathway unfeasible at room temperaturewithout large excess energy in the excitation pulse. Instead,the only major intramolecular nuclear relaxation pathwayinvolves subtle buckling of the xanthene core and slight rotationof the dialkylamino groups (Figure 2b, inset). From these TD-DFT calculations we deduce that the short lifetime componentin the TA of SRB is primarily a result of nuclear relaxationalong the intramolecular and solvent reaction coordinates,rather than internal conversion between different excited states.The spectral changes accompanying these processes are aredshift of the stimulated emission between 580 and 610 nmand a slight redshift of the excited state absorption (ESA) at450 nm, as seen more clearly in the decay-associated spectra(i.e., the spectral amplitudes of the exponential components ofthe global kinetic fits, Figure 2b). The population in the S1excited state then decays back to the ground state onnanosecond time scales.We observe several revealing trends in the rate of excited

state evolution that clearly demonstrate the importance ofchromophore−protein linker configuration and positioning inbiomimetic light-harvesting systems. Figure 2d and Table 1display the lifetimes associated with the picosecond solvationand intramolecular reorganization process for all systems. Thenanosecond relaxation component lifetimes are shown in Table1 and Figure S5 and follow approximately similar trends, albeit

less prominently. The most dominant features immediatelyobserved from Figure 2d are the up to 5-fold slowing ofsolvation and structural rearrangement when the chromophoresare attached to cpTMV at outer surface site S23C, and up to14-fold slowing when attached at intracavity site Q101C, whencompared to chromophores in buffer. The fact that dynamicretardation is clearly present even at the outside S23C site forthe more rigid linkers (all except butyl) suggests that thechromophores are close enough to the protein−solventinterface to be directly affected by it. Interestingly, however,SRB-Bu, which has the longest and most flexible linker, doesnot exhibit dynamic retardation at the outside position, insteaddisplaying the same time scales as for the free chromophore insolution. This result indicates that even if the chromophore iscovalently attached to the protein, provided the linker is longand flexible enough, it remains unaffected by the latter and onlysamples the bulk solvent environment far away from theprotein surface. Inside the protein cavity, SRB-Bu is sloweddown by an amount similar to other complexes. At thisattachment site, confinement by the top and bottom proteinsurfaces, separated by ∼2 nm,12 precludes the presence of abulk-like environment, leading to much more constraineddynamics irrespective of the linker used.Another surprising feature in Figure 2d is that the greatest

retardation is experienced by SRB-Cyc-SS both inside andoutside of the cavity. This difference is especially salient whencomparing the enantiomers Cyc-SS and Cyc-RR, with theformer slowing by a factor greater than the latter by 1.7 times atthe outside position and by 2.5 times inside the cavity. Thus, achiral modification of the chromophore linker translates intomajor changes in the excited state dynamics of the system. Thisfinding suggests that enantiomers can be selected for purposesof generating slow excited state dynamics that approximate theenvironmental protection of chromophore electronic excita-tions in natural photosynthesis. Overall, in line with ourprevious findings using generic linkers,12 the interior proteincavity can be used to constrain both chromophores and watermolecules and, in this way, substantially slow relaxation alongthe electronic excited state potential energy surface. Yet, goingbeyond our previous findings relating to confinement, we alsolearn here that purpose-specific linkers enable dynamicretardation at the protein surface (which would be absentusing the long tethers typical for commercial bioimaging dyes),and provide far better control over the extent to whichchromophore−protein interactions tune the excited statedynamics both inside and outside the protein cavity. Thesecompounds therefore facilitate a much more systematic and

Table 1. Short and Long Lifetime Components, along withDynamic Retardation Ratios Resulting from ChromophoreAttachment to the Protein at Different Sites, Extracted fromGlobal Analysis of TA Data for All Measured Systems

linkerτ

component buffer S23c Q101CS23c/buffer

Q101C/buffer

Butyl short (ps) 1.7 1.7 18 1.0 11long (ns) 1.7 1.9 2.8 1.1 1.6

Ethyl short (ps) 2.2 7.3 16 3.3 7.3long (ns) 1.5 2.1 2.4 1.4 1.6

Cyc-SS short (ps) 1.8 9.9 26 5.5 14long (ns) 1.4 2.1 2.3 1.5 1.6

Cyc-RR short (ps) 1.9 6.1 11 3.2 5.8long (ns) 1.9 1.9 3.1 1.0 1.6

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.7b13598J. Am. Chem. Soc. 2018, 140, 6278−6287

6281

Page 5: Exploiting Chromophore–Protein Interactions through Linker …geisslergroup.org/wp-content/uploads/2019/06/Pub2018-5.pdf · 2019-06-11 · shows the overlaid optimized ground and

precise way to study the role of chromophore−protein−solventinteractions in light-harvesting systems, as will be shown indetail below.To obtain a molecular-level understanding of the trends in

excited state dynamics observed for the four chromophore-linker systems attached to the surface site of cpTMV, weinvestigate how each linker alters the orientation andpositioning of the chromophore on the protein surface usingMD simulations. For computational tractability, a reducedsystem consisting of surface helices 1 and 2 of a cpTMVmonomer, along with helix 2 of an adjacent monomer issimulated. The chromophore is attached to S23C on helix 1. Aclose analogue of SRB, AlexaFluor 488, is used since accurateforce field parameters for AlexaFluor 488 have been developedto reproduce experimental data,38 whereas parameters have notyet been developed for SRB. The MD simulations not onlyconfirm an intuitive picture of cyclohexyl-linked chromophoresbeing constrained closer to the protein surface, but also reveal adetailed and subtle picture of chromophore−protein−waterinteractions that successfully explains the observed trends inexcited state dynamics.

More specifically, we consider how each linker determinesthe extent to which the chromophore is localized at the proteinsurface. The increased length and flexibility of the butyl andethyl linkers allow the chromophore to sample many differentconfigurations that reach far from the protein surface. Incontrast, the Cyc-RR and Cyc-SS constrain the chromophore inspecific configurations relative to the protein surface. In Figure3a, the distribution of the minimum distance between anycarbon atom of the chromophore’s xanthene core and a heavyprotein atom is plotted. For all four systems, the chromophorecan specifically interact with protein side chains, resulting in thepeak at ∼0.35 nm. The distributions for the butyl and ethylsystems exhibit a broad second peak centered at ∼1.2 nm and∼1 nm, respectively, whereas the distributions for thecyclohexyl systems both exhibit a second peak centered at∼0.7 nm. The 3D density plots of the xanthene heavy atoms ofthe chromophore (Figure 3b), averaged over the last 500 ns ofthe MD simulations, provide further insight into commonchromophore conformations on the protein surface. Plots at alow density isovalue (0.005 g/Å3) show all possible locationswhere the chromophore can be found during the simulations.Consistent with the distance distributions, the chromophore

Figure 3. Results of MD simulations on a cpTMV model system consisting of the chromophore attached to a cpTMV monomer (two surfacehelices), along with a single helix (at left) from an adjacent monomer. (a) Calculated chromophore−protein distance distribution functions.Minimum Euclidean distances between a xanthene carbon atom and protein heavy atom are used. (b) Side views of the densities of the chromophorexanthene atoms for each complex, along with a representative snapshot of the chromophore attachment position and orientation on the proteinsurface. In this displayed perspective, the two monomer helices are on the right, and the adjacent monomer’s helix is on the left. “N-term” indicatesthe N-terminus of helix 1, on which the chromophore is attached. Additional perspectives are shown in Figure S10. Density isosurfaces of thexanthene heavy atoms of the chromophore averaged over the final 500 ns of the MD simulations are shown for density values of 0.005 g/Å3 (lightershade) and 0.06 g/Å3 (darker shade), displaying much sparser distributions for butyl and ethyl compared to Cyc-SS and Cyc-RR-linkedchromophores. (c) Rotational time correlation function of the chromophore transition dipole moment. (d) Water dipole moment rotational timecorrelation function for bulk water (black line) vs water molecules within 5 Å of any xanthene heavy atom of the chromophore. Water mean squareddisplacement curves are displayed in Figure S11. (e) Diffusion coefficients of waters within 5 Å of each protein side chain (heavy atoms or Cα)plotted in false color scale in the absence (left, dashed box) and presence (right, Ethyl, Cyc-SS, Cyc-RR) of the chromophore. Additionalperspectives are shown in Figure S12. When the chromophore is present, water displacement is considerably slowed, particularly when thechromophore is linked through Cyc-SS, as the chromophore sandwiches water molecules between itself and a region of high protein side chaindensity (black arrow).

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.7b13598J. Am. Chem. Soc. 2018, 140, 6278−6287

6282

Page 6: Exploiting Chromophore–Protein Interactions through Linker …geisslergroup.org/wp-content/uploads/2019/06/Pub2018-5.pdf · 2019-06-11 · shows the overlaid optimized ground and

samples the largest expanse of space when attached via butyland ethyl linkers, as compared to Cyc-RR or Cyc-SS. Plots at amore stringent isovalue (0.06 g/Å3) help to identify the mostcommon conformations of the chromophore on the proteinsurface. Since the butyl and ethyl linkers are quite flexible, thebutyl- and ethyl-linked chromophores adopt many differentconformations and reside in multiple regions on the proteinsurface with densities less than 0.06 g/Å3. As a result, large,significant regions at this density isovalue are not observed. Incontrast, the Cyc-RR- and Cyc-SS-linked chromophores adoptonly a few, very specific conformations on the protein such thatsignificant regions of density are observed at the more stringentisovalue. These results support the conclusion that chromo-phores attached via either butyl or ethyl sample more regions ofspace farther from the protein surface than those attached viathe cyclohexyl groups, which remain within the biomolecularhydration shells.39

An additional consequence of restricting chromophoreconformational sampling via cyclohexyl linkers is the abilityto preserve orientational correlations for longer. In Figure 3c,the transition dipole moment orientational correlation functionof the chromophore is plotted for all systems, showing thatrotational reorientation is considerably slower for Cyc-SS- andCyc-RR-linked chromophores compared to ethyl- and butyl-linked complexes. The cyclohexyl linkers retain orientationalcorrelations over several nanoseconds, i.e., over the full excitedstate lifetimes of the chromophores. In light-harvestingassemblies, the ability to retain orientational correlationsbetween adjacent chromophores throughout the excitonlifetime allows for direct optimization of nonradiative energytransfer. Such control is necessary to reproduce the directionaland efficient energy transport over long distances that isachieved in natural systems. The cyclohexyl linkers aretherefore ideal chromophore−protein linkers for biomimeticlight-harvesting assemblies.To explain the differences between the excited state

dynamics of Cyc-RR and Cyc-SS-linked chromophores, weinvestigated the dynamics of water molecules around thechromophore and the protein surface. The molecularinterpretation of chromophore dynamic retardation due toslowed water molecules near biomolecular surfaces has beenstudied in depth by a wide range of experimental andcomputational techniques and remains somewhat controver-sial.39−48 Mounting evidence suggests that water dynamics nearprotein surfaces are not affected as much as previously thought,typically only slowing by a factor of 2−3 over the first hydrationshells, with the effect rapidly dropping off radially toward bulkwater.49,50 As a result, the commonly observed order-of-magnitude retardation in the excited state dynamics ofchromophores located in the first few hydration shells isoften attributed to slow biomolecular motion affecting bothwater and chromophore dynamics in its heterogeneousdielectric environment.50−55 Figure 3d displays the rotationaltime correlation function of the water dipole moments formolecules in the bulk (far away from the protein surface)compared to those within 5 Å of any of the chromophores’xanthene heavy atoms in our cpTMV system. Thesesimulations show that the water dynamics near thechromophores are slowed by a factor of ∼2, corroboratingthe aforementioned studies, but with the largest retardationoccurring around the Cyc-SS linker (see also Figure S11 andTable S1). The difference in water dynamics around Cyc-RR-and Cyc-SS-linked chromophores is a priori surprising given

that they reside a similar distance away from the proteinsurface. The dissimilarity suggests that a mechanism based onspecific chromophore−protein−solvent interactions is respon-sible for the observed dynamic differences, as discussed below.A closer look at the different chromophore orientations for

the Cyc-SS and Cyc-RR complexes in Figures 3b,e reveals thatthe chromophore attached via Cyc-SS resides primarily abovethe two helices of the same monomer, where a high density ofhydrophobic amino acid residues are located, whereas thechromophore attached via Cyc-RR resides above a less denseregion between adjacent monomers (Figure 3e). To determinea more detailed picture of slowed water dynamics, and whetherit can be attributed to the protein residues alone or to both theprotein and chromophore, the dynamics of waters within 5 Å ofeach protein side chain were analyzed for the model cpTMVsystem with and without a chromophore (Figures 3e and S12).The water diffusion coefficients are plotted in false color scalearound the protein surface, with blue representing the greatestretardation in water dynamics. Without the chromophorepresent (left), it is already apparent that water dynamics areslowest around densely packed areas of the protein, such as thearea between the helices of a cpTMV monomer, and withindeeper pockets or grooves along the protein surface. Strikingly,when the chromophore is present, the water dynamics aroundspecific residues are further slowed, suggesting that thechromophore also impacts the solvent dynamics. Thisretardation is particularly noticeable for the Cyc-SS complex,which is constrained just above a dense region of the proteinsurface. The same trends are observed when the dipolerotational correlation functions of waters around each residueare compared (Figure S13). These observations suggest thatwater molecules are effectively sandwiched between thechromophore and the protein surface, leading to retardationnear the surface that is a factor of ∼1.5 greater than in theabsence of the chromophore. For the ethyl complex, whichmainly resides in a similar but more distant region, watermolecules near the surface are slightly slowed by a factor of∼1.1 compared to without the chromophore. Finally, the Cyc-RR complex, which primarily resides above the more recessedand sparser groove between adjacent monomers, has negligibleeffect on water dynamics in the already fairly dynamicallyconstrained groove compared to the system without thechromophore.The above differences in water dynamics correlate with the

retardation trends observed experimentally between Ethyl-,Cyc-SS- and Cyc-RR-linked complexes at S23C (yellow bars inFigure 2d) remarkably well. In the TA, the intramolecular andsolvent dynamics for Et-, Cyc-SS, and Cyc-RR-linkedchromophores are slowed by factors of 3.3, 5.5, and 3.2,respectively (Table 1). In the simulations, the compoundedretardation of water molecules between the protein andchromophore are by factors of ∼4.8, 6.4, and 4.6, respectively.These results show that explicit inclusion of the chromophorein simulations is paramount to accurately determine the systemdynamics, and that in certain cases, the strong interactionbetween chromophore, protein and solvent leads to asupracomplex with intricately coupled dynamics. We thereforebelieve that a combination of slowed hydration dynamics andprotein motion, the latter being particularly effective atintracavity labeling sites where chromophores are surroundedby protein residues, are responsible for the up to 14-foldretardation (for Cyc-SS linked to inner cavity site Q101C) insolvent-chromophore dynamics in our experiments. By

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.7b13598J. Am. Chem. Soc. 2018, 140, 6278−6287

6283

Page 7: Exploiting Chromophore–Protein Interactions through Linker …geisslergroup.org/wp-content/uploads/2019/06/Pub2018-5.pdf · 2019-06-11 · shows the overlaid optimized ground and

engineering the distance from the protein surface and thevolumes sampled by chromophores in relation to specificprotein residues, solvent−solute interactions can be finely andpredictively tuned. The complex interplay between protein,water, and chromophore motions and dielectric interactionsthus provides an exceptionally rich and system-specific platformto control the excited state properties of biomimetic assemblies.

■ CONCLUSIONSToward developing biomimetic light harvesting assemblies, wehave combined a highly tunable protein scaffold based on thetobacco mosaic virus coat protein with a library ofchromophore−protein linkers. This combination enabled usto systematically study the molecular dynamics of chromo-phore−protein−solvent interactions and their effect on theultrafast excited state dynamics of the chromophores. Smallmodifications to the position, mobility, and attachmentorientation of the chromophores on the protein surface,achieved through alkyl chain shortening or switching the chiralconfigurations of the linkers, suffice to control the nuclearrelaxation dynamics of the system considerably. These resultstherefore highlight the often-overlooked importance ofpurpose-specific linker engineering in combination withjudicious attachment positioning to tune the properties ofartificial light-harvesting systems.On the basis of these findings, we describe below how our

assemblies enable a modular approach to tune the balancebetween chromophore−protein and interchromophore cou-plings, and how such fine-tuning will permit further under-standing of how nature has optimized its complex photo-synthetic environment for light harvesting. First, at the level ofchromophore−protein interactions, we found that minormodifications to chromophore−protein linker length andrigidity transform into pronounced differences in thechromophore properties when bound to the protein assembly.For example, shortening the butyl linker by just two bonds toethyl reins in the chromophore closer to the protein surface andinto a region prone to chromophore−protein interactions. Thisshortening changes the chromophore environment from beingcompletely bulk-like to the more heterogeneous protein−solvent interface. Consequently, a greater than 4-fold slowing ofthe Stokes shift time scale occurs at the outside labelingposition between butyl-linked and ethyl-linked chromophores.Further rigidifying the chromophore to cyclohexyl-based linkersenhances the retardation up to 6-fold. This strategy to tune theoptical properties of chromophores and preserve electronicexcitations in the system by coupling the chromophore andprotein may be a crucial photosynthetic design principle, wherechromophores are tightly bound within protein pockets andthus necessarily interact with their biomolecular environment.These considerations invite a more explicit comparison of

our biomimetic light harvesting platform with the natural onesthat it seeks to emulate. In light-harvesting systems whereinterchromophore energy transfer between identical chromo-phores is present, retarding Stokes shift dynamics could enablea larger number of resonant energy hops to proceed well out-of-equilibrium within the lifetime of an excitation. These energyhops can be more efficient than those initiated from structurallyrelaxed chromophores due to the larger spectral overlapbetween adjacent chromophores prior to nuclear relaxation,thus aiding long-range energy transport. Furthermore, thepreservation of excess energy within the system can helpovercome unwanted trapping in low-energy states that are

inevitable in disordered environments. While individualchromophore dynamics in natural light-harvesting systems areunknown due to our inability to study them in the absence ofinterchromophore coupling, their degree of confinement, rangeof motion and distance from protein surfaces in complexes suchas the LH2 of purple bacteria or the FMO of green sulfurbacteria are on the same order as for our cyclohexyl-basedcomplexes.22,56 Chromophores in these natural systems arethus likely subject to similar or greater dynamic retardation oftheir individual Stokes shift dynamics. Our highly tunablescaffold will facilitate a bottom-up approach to test ourhypotheses on the role of relaxation retardation on light-harvesting and provide direct insight into whether thesemechanisms are also likely to be operative in photosyntheticorganisms.Second, looking toward the control of interchromophore

interactions, we have shown that further constraints onchromophore environments can be readily applied by lockingthe chromophore into specific positions on the protein surface.To that end, we found that linker chirality can be exploited:despite no major difference in length or rigidity, changing thelinker from (R,R)- to (S,S)-cyclohexyl leads to considerablyslower excited state dynamics due to different attachmentpositions. These different configurations lead to distinctchromophore−protein bath interactions, including sandwichingwater molecules in a tight space between chromophore andprotein. Furthermore, both cyclohexyl-linked chromophoresexhibit much longer orientational decorrelation times than themore flexible alkyl linkers. Over the full course of thechromophores’ excited state lifetimes, this constraint enablesgreater control of the relative orientations of the chromophoreswith respect to each other and the protein surfacea crucialfeature to achieve long-range energy transport. The ability tomanipulate the relative position and orientation of adjacentchromophores in easily configurable artificial light harvestingconstructs presents the further possibility to mimic and test thecharacteristics of specific interchromophore motifs found innature.Having refined chromophore−protein interactions through

our studies performed in the absence of interchromophoreinteractions, we are now poised to reintroduce theseinteractions via saturating labeling densities to form completebiomimetic light harvesters. A key challenge will be to obtainthe right balance between interchromophore and chromo-phore−protein coupling for a range of energy transport regimesfound in nature, from long-range dipole−dipole to short-rangeexcitonic processes, while maintaining systems free of contactquenching. The degree of control over chromophore−proteininteractions that we have demonstrated in our bottom-upapproach, and the flexibility over attachment positioning andhence interchromophore distance afforded by our syntheticscaffold, will prove essential in finding the tuning range overwhich energy transfer proceeds efficiently for different transportregimes. Examining the role of chromophore−solvent−proteininteractions over the entirety of this tuning range will enabletesting and refining our current models seeking to explain thehigh quantum efficiencies and vast diversity of photosyntheticorganisms. To mimic nature even further, the construction ofextended arrays of cpTMV rings, analogous to the arrays ofnatural light harvesting complexes packed into lipid mem-branes, could allow the exploration of longer range biomimeticenergy transfer using recently developed spatially resolvedapproaches.57

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.7b13598J. Am. Chem. Soc. 2018, 140, 6278−6287

6284

Page 8: Exploiting Chromophore–Protein Interactions through Linker …geisslergroup.org/wp-content/uploads/2019/06/Pub2018-5.pdf · 2019-06-11 · shows the overlaid optimized ground and

Ultimately, the fine level of control over linker propertiesachievable through well-known synthetic procedures, alongwith high-throughput screening of candidate linkers usingcomputational modeling, represents a highly effective strategyto design purpose-specific biomimetic tools to more easily testhypotheses regarding the molecular mechanisms of naturalphotosynthetic light harvesting. In particular, since nuclear-electronic coupling often dictates the fate of molecular excitedstates following light absorption, harnessing chromophore−bath couplings to affect the structural motion of electronicallycoupled arrays of chromophores58−61 could prove paramountto achieving long-range energy transport as efficiently as innatural photosynthetic systems. Such an achievement wouldhelp elucidate the molecular and intermolecular origin of theunparalleled efficiency of photosynthetic light harvesting.

■ METHODSSynthesis. Extensive details of the synthetic procedures, mutant

generation, protein expression and bioconjugation and purity analysesare provided in the Supporting Information. The collection ofmaleimide functionalized sulforhodamine B dyes spaced by variouslinkers were synthesized starting from commercially available alcohol-amines. The Boc-protection was first selectively carried out on theamine group, followed by Mitsunobu displacement of the free alcoholwith maleimide. After released by deprotection, the free amine wasable to couple the sulforhodamine B acid chloride to yield therespective sulforhodamine B maleimides.62 The statistically singlelabeling of the double disk protein assembly was achieved by adjustingthe equivalents of the functionalized dyes and the cysteine containingprotein monomers as low as 0.01. The labeling yields of the proteinbioconjugates were analyzed by a Time-of-Flight (TOF) massspectrometer with a dual electrospray source connected in-line withhigh-performance liquid chromatography (ESI-TOF LC−MS).Analytical size exclusion chromatography was performed for assess-ment of purity and validation of assembly state.Transient Absorption. Ultrafast transient absorption spectrosco-

py was performed using a Ti:sapphire regenerative amplifier delivering5 W, 5 kHz, 40 fs pulses centered at 800 nm. Pump pulses centered at570 nm with 40 fs pulsewidth were generated in a home-buildnoncollinear parametric amplifier followed by a prism-pair compressor.The probe was a white light supercontinuum generated by focusing800 nm light in a CaF2 crystal. Pump and probe are set at magic angle(54.7°) polarization to each other. The pump power density used was∼300 μJ/cm2, well within a linear excitation regime. Optical densitiesat the excitation wavelength were at ∼0.1. All experiments areperformed at room temperature. The samples were continuouslyrastered during data acquisition, and UV−vis spectra taken before andafter every run verified that no sample decomposition occurred.DFT. Vertical excitations energies, oscillator strengths, and natural

transition orbitals (NTOs) with ground (S0) and excited state (S1)minimum energy geometries were calculated in Gaussian63 utilizingTDDFT at the PBE064/6-31+g*65 level, with aqueous solvation effectsfrom a standard polarizable continuum model utilizing the integralequation formalism (IEF-PCM).66 The effect of the exchangecorrelation functional (XCF) on the excitation energies was checkedby comparing PBE67 (a generalized gradiant approximation functional)two global hybrids (PBE064 and B3LYP68,69) and two range-separatedhybrids (CAM-B3LYP70 and ωB97X-D71,72). As the Hartree−Fock(HF) exchange included in a functional increases, the S1 and S2excitation energies also increase, as do the energy differences betweenthe two states. However, the qualitative picture remains the same, withorbital densities and excitation character similar between functionals(Figure S7).MD Simulations. Simulation methods are briefly described below,

and complete details are given in the Supporting Information. Toinvestigate the protein environment around each chromophore, weperformed MD simulations of a system composed of three surfaceexposed α-helices of cpTMV from the crystal structure of cpTMV

(PDB: 3KML), in which the chromophore was linked to S23C on thecentral α-helix via each organic linker. Because force field parametersfor sulforhodamine B have not been published, a similar rhodaminedye, AlexaFluor 488, was simulated. We used the Amber03ws forcefield73 and TIP4P/2005s water model73,74 since the parameters ofAlexaFluor 488 have been optimized to reproduce experimental datausing this combination of protein force field and water model.38 Toefficiently sample configurations of the chromophore on the proteinsurface, 1 μs bias-exchange metadynamics (BE-META) simulations75

were performed using GROMACS 4.676 with the PLUMED 2 patch.77

For all four systems with a chromophore, a bias was applied on theS23C dihedrals χ1 (N−Cα−Cβ−Sγ), χ2 (Cα−Cβ−Sγ−Cmal, whereCmal is the carbon of the maleimide group attached to Cys). Additionalbiases were applied to enhance sampling of all rotatable bonds of thetwo aliphatic linkers: For butyl, three additional biases were applied onthe dihedrals Nmal−C1−C2−C3, C1−C2−C3−C4, and C2−C3−C4−Namide, where Nmal is the nitrogen of the maleimide group, C1, C2, C3,and C4 are the carbons of the butyl linker, and Namide is the nitrogen ofthe amide bond connecting the linker to the chromophore. For ethyl,one additional bias was applied on the dihedral Nmal−C1−C2−Namide,where C1 and C2 are the carbons of the ethyl linker. For all systems,one unbiased replica was also included, and the last 500 ns of thisreplica was used for accurate analysis of the unbiased ensemble. Toinvestigate the dynamics of the chromophore and solvent,configurations obtained from the last 500 ns of the BE-METAsimulations were used to initialize ten short MD simulations for eachsystem. For comparison, short MD simulations of a system without achromophore were also performed.

■ ASSOCIATED CONTENT*S Supporting InformationThe Supporting Information is available free of charge on theACS Publications website at DOI: 10.1021/jacs.7b13598.

Detailed materials and methods; all transient absorptionspectra and associated global analysis and lifetimes;additional results from DFT calculations on SRB,including ground and excited state optimized geometries,details on exchange correlation functional effects oncalculated excitation energies, and natural transitionorbitals; additional results from MD simulations,including water diffusion coefficients and additionalperspectives of 3D density plots (PDF)

■ AUTHOR INFORMATIONCorresponding Author*[email protected] Delor: 0000-0002-9480-9235Matthew B. Francis: 0000-0003-2837-2538Naomi S. Ginsberg: 0000-0002-5660-3586Author Contributions⊗M.D., J.D., T.D.R., and J.R.R. contributed equally to thiswork.NotesThe authors declare no competing financial interest.

■ ACKNOWLEDGMENTSThis work has been supported by the Director, Office ofScience, Chemical Sciences, Geosciences, and BiosciencesDivision, of the U.S. Department of Energy under ContractNo. DEAC02-05CH11231. T.D.R. acknowledges a NationalScience Foundation Graduate Research Fellowship (DGE1106400). N.S.G. acknowledges an Alfred P. Sloan ResearchFellowship, a David and Lucile Packard Foundation Fellowship

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.7b13598J. Am. Chem. Soc. 2018, 140, 6278−6287

6285

Page 9: Exploiting Chromophore–Protein Interactions through Linker …geisslergroup.org/wp-content/uploads/2019/06/Pub2018-5.pdf · 2019-06-11 · shows the overlaid optimized ground and

for Science and Engineering, a Camille and Henry DreyfusTeacher-Scholar Award, and support of the Miller Institute forBasic Research in Science.

■ REFERENCES(1) Blankenship, R. E. Molecular Mechanisms of Photosynthesis;Blackwell Science: Oxford, U.K., 2002.(2) Cheng, Y.-C.; Fleming, G. R. Annu. Rev. Phys. Chem. 2009, 60,241−262.(3) Scholes, G. D.; Fleming, G. R.; Olaya-Castro, A.; van Grondelle,R. Nat. Chem. 2011, 3, 763−774.(4) Croce, R.; van Amerongen, H. Nat. Chem. Biol. 2014, 10, 492−501.(5) Nam, Y. S.; Shin, T.; Park, H.; Magyar, A. P.; Choi, K.; Fantner,G.; Nelson, K. A.; Belcher, A. M. J. Am. Chem. Soc. 2010, 132, 1462−1463.(6) Springer, J. W.; Parkes-Loach, P. S.; Reddy, K. R.; Krayer, M.;Jiao, J.; Lee, G. M.; Niedzwiedzki, D. M.; Harris, M. A.; Kirmaier, C.;Bocian, D. F.; Lindsey, J. S.; Holten, D.; Loach, P. A. J. Am. Chem. Soc.2012, 134, 4589−4599.(7) Cohen-Ofri, I.; van Gastel, M.; Grzyb, J.; Brandis, A.; Pinkas, I.;Lubitz, W.; Noy, D. J. Am. Chem. Soc. 2011, 133, 9526−9535.(8) Farid, T. A.; Kodali, G.; Solomon, L. A.; Lichtenstein, B. R.;Sheehan, M. M.; Fry, B. A.; Bialas, C.; Ennist, N. M.; Siedlecki, J. A.;Zhao, Z.; Stetz, M. A.; Valentine, K. G.; Anderson, J. L. R.; Wand, A. J.;Discher, B. M.; Moser, C. C.; Dutton, P. L. Nat. Chem. Biol. 2013, 9,826−833.(9) Endo, M.; Fujitsuka, M.; Majima, T. Chem. - Eur. J. 2007, 13,8660−8666.(10) Miller, R. A.; Presley, A. D.; Francis, M. B. J. Am. Chem. Soc.2007, 129, 3104−3109.(11) Dedeo, M. T.; Duderstadt, K. E.; Berger, J. M.; Francis, M. B.Nano Lett. 2010, 10, 181−186.(12) Noriega, R.; Finley, D. T.; Haberstroh, J.; Geissler, P. L.;Francis, M. B.; Ginsberg, N. S. J. Phys. Chem. B 2015, 119, 6963−6973.(13) Peng, H.-Q.; Niu, L.-Y.; Chen, Y.-Z.; Wu, L.-Z.; Tung, C.-H.;Yang, Q.-Z. Chem. Rev. 2015, 115, 7502−7542.(14) Gust, D.; Moore, T. A.; Moore, A. L. Acc. Chem. Res. 2001, 34,40−48.(15) Utschig, L. M.; Soltau, S. R.; Tiede, D. M. Curr. Opin. Chem.Biol. 2015, 25, 1−8.(16) Renger, T. Photosynth. Res. 2009, 102, 471−485.(17) Lee, H.; Cheng, Y.-C.; Fleming, G. R. Science 2007, 316, 1462−1465.(18) Arpino, J. A. J.; Czapinska, H.; Piasecka, A.; Edwards, W. R.;Barker, P.; Gajda, M. J.; Bochtler, M.; Jones, D. D. J. Am. Chem. Soc.2012, 134, 13632−13640.(19) Sarovar, M.; Birgitta Whaley, K. New J. Phys. 2013, 15, 13030.(20) van Grondelle, R.; Novoderezhkin, V. I. Phys. Chem. Chem. Phys.2006, 8, 793−807.(21) Renger, T.; May, V.; Kuhn, O. Phys. Rep. 2001, 343, 137−254.(22) McDermott, G.; Prince, S. M.; Freer, A. A.; Hawthornthwaite-Lawless, A. M.; Papiz, M. Z.; Cogdell, R. J.; Isaacs, N. W. Nature 1995,374, 517−521.(23) Jordan, P.; Fromme, P.; Witt, H. T.; Klukas, O.; Saenger, W.;Krauss, N. Nature 2001, 411, 909−917.(24) Liu, Z.; Yan, H.; Wang, K.; Kuang, T.; Zhang, J.; Gui, L.; An, X.;Chang, W. Nature 2004, 428, 287−292.(25) Hu, X.; Ritz, T.; Damjanovic, A.; Autenrieth, F.; Schulten, K. Q.Rev. Biophys. 2002, 35, 1−62.(26) Nam, Y. S.; Shin, T.; Park, H.; Magyar, A. P.; Choi, K.; Fantner,G.; Nelson, K. A.; Belcher, A. M. J. Am. Chem. Soc. 2010, 132, 1462−1463.(27) Springer, J. W.; Parkes-Loach, P. S.; Reddy, K. R.; Krayer, M.;Jiao, J.; Lee, G. M.; Niedzwiedzki, D. M.; Harris, M. A.; Kirmaier, C.;Bocian, D. F.; Lindsey, J. S.; Holten, D.; Loach, P. A. J. Am. Chem. Soc.2012, 134, 4589−4599.

(28) Ma, Y.-Z.; Miller, R. A.; Fleming, G. R.; Francis, M. B. J. Phys.Chem. B 2008, 112, 6887−6892.(29) Wen, A. M.; Steinmetz, N. F. Chem. Soc. Rev. 2016, 45, 4074−4126.(30) Sabatini, R. P.; Mark, M. F.; Mark, D. J.; Kryman, M. W.; Hill, J.E.; Brennessel, W. W.; Detty, M. R.; Eisenberg, R.; McCamant, D. W.Photochem. Photobiol. Sci. 2016, 15, 1417−1432.(31) Grabowski, Z. R.; Rotkiewicz, K.; Rettig, W. Chem. Rev. 2003,103, 3899−4031.(32) Savarese, M.; Aliberti, A.; De Santo, I.; Battista, E.; Causa, F.;Netti, P. A.; Rega, N. J. Phys. Chem. A 2012, 116, 7491−7497.(33) Savarese, M.; Raucci, U.; Adamo, C.; Netti, P. A.; Ciofini, I.;Rega, N. Phys. Chem. Chem. Phys. 2014, 16, 20681−20688.(34) Zhang, X.-F.; Zhang, Y.; Liu, L. J. Lumin. 2014, 145, 448−453.(35) Fedoseeva, M.; Letrun, R.; Vauthey, E. J. Phys. Chem. B 2014,118, 5184−5193.(36) Grabowski, Z. R.; Rotkiewicz, K.; Rettig, W. Chem. Rev. 2003,103, 3899−4032.(37) Savarese, M.; Raucci, U.; Adamo, C.; Netti, P. A.; Ciofini, I.;Rega, N. Phys. Chem. Chem. Phys. 2014, 16, 20681−20688.(38) Best, R. B.; Hofmann, H.; Nettels, D.; Schuler, B. Biophys. J.2015, 108, 2721−2731.(39) Laage, D.; Elsaesser, T.; Hynes, J. T. Chem. Rev. 2017, 117,10694−10725.(40) Jordanides, X. J.; Lang, M. J.; Song, X.; Fleming, G. R. J. Phys.Chem. B 1999, 103, 7995−8005.(41) Pal, S. K.; Zewail, A. H. Chem. Rev. 2004, 104, 2099−2124.(42) Bagchi, B. Chem. Rev. 2005, 105, 3197−3219.(43) Nilsson, L.; Halle, B. Proc. Natl. Acad. Sci. U. S. A. 2005, 102,13867−13872.(44) Li, T.; Hassanali, A. A.; Kao, Y.-T.; Zhong, D.; Singer, S. J. J. Am.Chem. Soc. 2007, 129, 3376−3382.(45) Furse, K. E.; Corcelli, S. A. J. Am. Chem. Soc. 2008, 130, 13103−13109.(46) Yang, M.; Szyc, L.; Elsaesser, T. J. Phys. Chem. B 2011, 115,13093−13100.(47) Sterpone, F.; Stirnemann, G.; Laage, D. J. Am. Chem. Soc. 2012,134, 4116−4119.(48) King, J. T.; Arthur, E. J.; Brooks, C. L.; Kubarych, K. J. J. Phys.Chem. B 2012, 116, 5604−5611.(49) Sterpone, F.; Stirnemann, G.; Laage, D. J. Am. Chem. Soc. 2012,134, 4116−4119.(50) Laage, D.; Elsaesser, T.; Hynes, J. T. Chem. Rev. 2017, 117,10694−10725.(51) Halle, B.; Nilsson, L. J. Phys. Chem. B 2009, 113, 8210−8213.(52) Furse, K. E.; Corcelli, S. A. J. Am. Chem. Soc. 2011, 133, 720−723.(53) Furse, K. E.; Corcelli, S. A. J. Am. Chem. Soc. 2008, 130, 13103−13109.(54) Homsi Brandeburgo, W.; van der Post, S. T.; Meijer, E. J.;Ensing, B. Phys. Chem. Chem. Phys. 2015, 17, 24968−24977.(55) Bakulin, A. A.; Pshenichnikov, M. S.; Bakker, H. J.; Petersen, C.J. Phys. Chem. A 2011, 115, 1821−1829.(56) Tronrud, D. E.; Wen, J.; Gay, L.; Blankenship, R. E. Photosynth.Res. 2009, 100, 79−87.(57) Penwell, S. B.; Ginsberg, L. D. S.; Noriega, R.; Ginsberg, N. S.Nat. Mater. 2017, 16, 1136−1142.(58) Wahadoszamen, M.; Margalit, I.; Ara, A. M.; van Grondelle, R.;Noy, D. Nat. Commun. 2014, 5, 5287.(59) Scholes, G. D.; Fleming, G. R.; Chen, L. X.; Aspuru-Guzik, A.;Buchleitner, A.; Coker, D. F.; Engel, G. S.; van Grondelle, R.; Ishizaki,A.; Jonas, D. M.; Lundeen, J. S.; McCusker, J. K.; Mukamel, S.; Ogilvie,J. P.; Olaya-Castro, A.; Ratner, M. A.; Spano, F. C.; Whaley, K. B.;Zhu, X. Nature 2017, 543, 647−656.(60) Fuller, F. D.; Pan, J.; Gelzinis, A.; Butkus, V.; Senlik, S. S.;Wilcox, D. E.; Yocum, C. F.; Valkunas, L.; Abramavicius, D.; Ogilvie, J.P. Nat. Chem. 2014, 6, 1−6.(61) Ghosh, S.; Bishop, M. M.; Roscioli, J. D.; LaFountain, A. M.;Frank, H. A.; Beck, W. F. J. Phys. Chem. Lett. 2017, 8, 463−469.

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.7b13598J. Am. Chem. Soc. 2018, 140, 6278−6287

6286

Page 10: Exploiting Chromophore–Protein Interactions through Linker …geisslergroup.org/wp-content/uploads/2019/06/Pub2018-5.pdf · 2019-06-11 · shows the overlaid optimized ground and

(62) Elduque, X.; Sanchez, A.; Sharma, K.; Pedroso, E.; Grandas, A.Bioconjugate Chem. 2013, 24, 832−839.(63) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; Li, X.; Caricato, M.; Marenich, A. V.; Bloino, J.;Janesko, B. G.; Gomperts, R.; Mennucci, B.; Hratchian, H. P.; Ortiz, J.V.; Izmaylov, A. F.; Sonnenberg, J. L.; Williams-Young, D.; Ding, F.;Lipparini, F.; Egidi, F.; Goings, J.; Peng, B.; Petrone, A.; Henderson,T.; Ranasinghe, D.; Zakrzewski, V. G.; Gao, J.; Rega, N.; Zheng, G.;Liang, W.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.;Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.;Throssell, K.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.;Bearpark, M. J.; Heyd, J. J.; Brothers, E. N.; Kudin, K. N.; Staroverov,V. N.; Keith, T. A.; Kobayashi, R.; Normand, J.; Raghavachari, K.;Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.;Millam, J. M.; Klene, M.; Adamo, C.; Cammi, R.; Ochterski, J. W.;Martin, R. L.; Morokuma, K.; Farkas, O.; Foresman, J. B.; Fox, D. J.Gaussian 09, Revision D.01; Gaussian, Inc.: Wallingford, CT, 2013.(64) Adamo, C.; Barone, V. J. Chem. Phys. 1999, 110, 6158−6170.(65) Ditchfield, R.; Hehre, W. J.; Pople, J. A. J. Chem. Phys. 1971, 54,724−728.(66) Tomasi, J.; Mennucci, B.; Cammi, R. Chem. Rev. 2005, 105,2999−3094.(67) Perdew, J. P.; Burke, K.; Ernzerhof, M. Phys. Rev. Lett. 1996, 77,3865−3868.(68) Becke, A. D. J. Chem. Phys. 1993, 98, 5648−5652.(69) Stephens, P. J.; Devlin, F. J.; Chabolowski, C. F.; Frisch, M. J. J.Phys. Chem. 1994, 98, 11623−11627.(70) Yanai, T.; Tew, D.; Handy, N. Chem. Phys. Lett. 2004, 393, 51−57.(71) Chai, J.-D.; Head-Gordon, M. J. Chem. Phys. 2008, 128, 84106.(72) Chai, J.-D.; Head-Gordon, M. Phys. Chem. Chem. Phys. 2008, 10,6615−6620.(73) Best, R. B.; Zheng, W.; Mittal, J. J. Chem. Theory Comput. 2014,10, 5113−5124.(74) Abascal, J. L. F.; Vega, C. J. Chem. Phys. 2005, 123, 234505.(75) Piana, S.; Laio, A. J. Phys. Chem. B 2007, 111, 4553−4559.(76) Hess, B.; Kutzner, C.; van der Spoel, D.; Lindahl, E. J. Chem.Theory Comput. 2008, 4, 435−447.(77) Tribello, G.; Bonomi, M.; Branduardi, D.; Camilloni, C.; Bussi,G. Comput. Phys. Commun. 2014, 185, 604−613.

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.7b13598J. Am. Chem. Soc. 2018, 140, 6278−6287

6287