tracking of the molecular motion in the primary event of...

11
Tracking of the Molecular Motion in the Primary Event of Photoinduced Reactions of a Phytochromobilin Model Xuhui Zhuang, ,,§ Jun Wang, ,,§ and Zhenggang Lan* ,,,§ Key Laboratory of Biobased Materials, Qingdao Institute of Bioenergy and Bioprocess Technology, Chinese Academy of Sciences, Qingdao, Shandong 266101, P. R. China University of Chinese Academy of Sciences, Beijing 100049, P. R. China § The Qingdao Key Lab of Solar Energy Utilization and Energy Storage Technology, Qingdao Institute of Bioenergy and Bioprocess Technology, Chinese Academy of Sciences, Qingdao, Shandong 266101, P. R. China * S Supporting Information ABSTRACT: The photoinduced processes of phytochromes have received great research interest for their important biological functions. Phytochromobilin (PΦB), one of the most important phytochrome chromophores, was selected as the prototype to study its photoinduced isomerization. The nonadiabatic dynamics of PΦB from the Pr conguration in the gas phase was investigated by the surface hopping method at the OM2/MRCI level. In the excited state, isolated PΦB does not display the rotation of the two terminal ve-membered rings (ring A and ring D), which is assumed to govern the Pr Pfr process in the protein. Instead, two S 1 /S 0 conical intersection seams (CI 01α and CI 01β ) characterized by the rotation of the two central rings (ring B and ring C) were proven to play essential roles for the photoisomerization of PΦB in the gas phase. These two conical intersections (CI 01α and CI 01β ) are accessible by the twisting motions of the C9C10 and C10C11 bonds, respectively. The CI 01α and CI 01β seams, instead of their minimum-energy points, are responsible for the nonadiabatic dynamics. For both channels, the trajectories may propagate forward to the isomerization products or backward to the original Pr conguration after the S 1 S 0 hops. 1. INTRODUCTION Phytochromes represent a widely existing group of important photoreceptors in plants, bacteria, and fungi. These photo- receptors can serve as bioclocks to regulate the fundamental functions of plants and other autotrophs, such as seed germination, growth, phototaxis, pigmentation, and so on. 1 Thus, some important phytochromes have been investigated frequently in the last decades. 14 From these studies, it is well- known that the functionality of phytochromes is governed by switching between two photoconvertible forms, namely, the physiologically inactive red form (Pr) and the physiologically active far-red form (Pfr), which show maximum photo- absorptions in the red (660 nm) and far-red (730 nm) regions, respectively. 1 The photoreaction centers of phyto- chromes were also identied and are composed of an open- chain terrapyrrole chromophore, such as phytochromobilin (PΦB), phycocyanobilin (PCB), or biliverdin (BV), sur- rounded by protein residues (Figure 1). Besides photoreceptors, the open-chain tetrapyrrole chro- mophores also play essential roles in light-harvesting systems and energy-transfer intermediates in antenna complexes. 4 Moreover, these chromophores are analogues of porphyrins, which represent another type of photoactive compounds with biological importance. The porphyrins and the phytochrome chromophores can be biosynthetically converted via ring opening reactions or vice versa. 4 The phytochrome chromophores have many isomers because of the existence of many twistable single and double CC bonds in such open-chain tetrapyrrole systems. This makes identifying their congurations a very challenging problem. As discussed in previous studies, 1 the Z/E isomerization happens at the double bonds of C4C5, C10C11, and C15C16, while the symmetry/antisymmetry (s/a) isomerization exists at single bonds such as C5C6, C9C10, and C14C15. By using this rule, we can easily dene the conformation of any isomer. In last decades, many pioneering works were devoted to determining the molecular congurations of the Pr conforma- tion of dierent phytochrome chromophores. 1 Among them, the Pr conformations of BV (in bacteria) and PCB (in cyanobacteria) were identied as the ZsZsZa stereoisomer (C5- Z,syn C10-Z,syn C15-Z,anti; see Figure 1) by crystallography and NMR spectroscopy, 510 while the Pr conformation of PΦB (in plant) was assigned as the ZaZsZa stereoisomer (C5-Z,anti C10-Z,syn C15-Z,anti; see Figure 2) by resonance Raman (RR) spectroscopy. 11,12 In constrast to the well-known conguration of the Pr form, the structure of the Pfr conformation seems to be controversial and is assumed to depend on the reaction mechanisms of the photoinduced Pr Pfr conversion processes in dierent Received: September 2, 2013 Revised: November 20, 2013 Published: November 21, 2013 Article pubs.acs.org/JPCB © 2013 American Chemical Society 15976 dx.doi.org/10.1021/jp408799b | J. Phys. Chem. B 2013, 117, 1597615986

Upload: others

Post on 21-Oct-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

  • Tracking of the Molecular Motion in the Primary Event ofPhotoinduced Reactions of a Phytochromobilin ModelXuhui Zhuang,†,‡,§ Jun Wang,†,‡,§ and Zhenggang Lan*,†,‡,§

    †Key Laboratory of Biobased Materials, Qingdao Institute of Bioenergy and Bioprocess Technology, Chinese Academy of Sciences,Qingdao, Shandong 266101, P. R. China‡University of Chinese Academy of Sciences, Beijing 100049, P. R. China§The Qingdao Key Lab of Solar Energy Utilization and Energy Storage Technology, Qingdao Institute of Bioenergy and BioprocessTechnology, Chinese Academy of Sciences, Qingdao, Shandong 266101, P. R. China

    *S Supporting Information

    ABSTRACT: The photoinduced processes of phytochromes have received great research interestfor their important biological functions. Phytochromobilin (PΦB), one of the most importantphytochrome chromophores, was selected as the prototype to study its photoinduced isomerization.The nonadiabatic dynamics of PΦB from the Pr configuration in the gas phase was investigated bythe surface hopping method at the OM2/MRCI level. In the excited state, isolated PΦB does notdisplay the rotation of the two terminal five-membered rings (ring A and ring D), which is assumedto govern the Pr → Pfr process in the protein. Instead, two S1/S0 conical intersection seams (CI01αand CI01β) characterized by the rotation of the two central rings (ring B and ring C) were proven toplay essential roles for the photoisomerization of PΦB in the gas phase. These two conicalintersections (CI01α and CI01β) are accessible by the twisting motions of the C9−C10 and C10−C11bonds, respectively. The CI01α and CI01β seams, instead of their minimum-energy points, areresponsible for the nonadiabatic dynamics. For both channels, the trajectories may propagateforward to the isomerization products or backward to the original Pr configuration after the S1 → S0hops.

    1. INTRODUCTION

    Phytochromes represent a widely existing group of importantphotoreceptors in plants, bacteria, and fungi. These photo-receptors can serve as bioclocks to regulate the fundamentalfunctions of plants and other autotrophs, such as seedgermination, growth, phototaxis, pigmentation, and so on.1

    Thus, some important phytochromes have been investigatedfrequently in the last decades.1−4 From these studies, it is well-known that the functionality of phytochromes is governed byswitching between two photoconvertible forms, namely, thephysiologically inactive red form (Pr) and the physiologicallyactive far-red form (Pfr), which show maximum photo-absorptions in the red (∼660 nm) and far-red (∼730 nm)regions, respectively.1 The photoreaction centers of phyto-chromes were also identified and are composed of an open-chain terrapyrrole chromophore, such as phytochromobilin(PΦB), phycocyanobilin (PCB), or biliverdin (BV), sur-rounded by protein residues (Figure 1).Besides photoreceptors, the open-chain tetrapyrrole chro-

    mophores also play essential roles in light-harvesting systemsand energy-transfer intermediates in antenna complexes.4

    Moreover, these chromophores are analogues of porphyrins,which represent another type of photoactive compounds withbiological importance. The porphyrins and the phytochromechromophores can be biosynthetically converted via ringopening reactions or vice versa.4

    The phytochrome chromophores have many isomers becauseof the existence of many twistable single and double C−Cbonds in such open-chain tetrapyrrole systems. This makesidentifying their configurations a very challenging problem. Asdiscussed in previous studies,1 the Z/E isomerization happensat the double bonds of C4−C5, C10−C11, and C15−C16,while the symmetry/antisymmetry (s/a) isomerization exists atsingle bonds such as C5−C6, C9−C10, and C14−C15. Byusing this rule, we can easily define the conformation of anyisomer. In last decades, many pioneering works were devoted todetermining the molecular configurations of the Pr conforma-tion of different phytochrome chromophores.1 Among them,the Pr conformations of BV (in bacteria) and PCB (incyanobacteria) were identified as the ZsZsZa stereoisomer (C5-Z,syn C10-Z,syn C15-Z,anti; see Figure 1) by crystallographyand NMR spectroscopy,5−10 while the Pr conformation of PΦB(in plant) was assigned as the ZaZsZa stereoisomer (C5-Z,antiC10-Z,syn C15-Z,anti; see Figure 2) by resonance Raman (RR)spectroscopy.11,12

    In constrast to the well-known configuration of the Pr form,the structure of the Pfr conformation seems to be controversialand is assumed to depend on the reaction mechanisms of thephotoinduced Pr → Pfr conversion processes in different

    Received: September 2, 2013Revised: November 20, 2013Published: November 21, 2013

    Article

    pubs.acs.org/JPCB

    © 2013 American Chemical Society 15976 dx.doi.org/10.1021/jp408799b | J. Phys. Chem. B 2013, 117, 15976−15986

    pubs.acs.org/JPCB

  • biological systems. Therefore, the study of photoinduced Pr →Pfr reactions becomes a very important topic not only for theunderstanding of the photoisomerization processes but also forthe assignment of the resulting Pfr structures.The Pr → Pfr conversions of phytochrome chromophores in

    protein environments have been widely investigated exper-imentally. Although the mechanism of photoinduced Pr → Pfrisomerization of phytochromes is still controversial,13 manystudies have agreed that the photoisomerization includes tworeaction steps. In step 1, the Z → E photoisomerization at theC15−C16 double bond plays the key role, and this step givesthe so-called “Lumi-R” intermediate.11,14−19 In step 2, the“Lumi-R” intermediate can convert to the Pfr isomer viathermal reactions (internal rotation around the C14−C15bond).15,16,18,19 But some researchers believed that the secondstep (the thermal one) may involve the twisting motion of theC5−C6 bond between ring A and ring B.11,12 Since it is stillvery difficult to get more detailed information, some scientistshave tried to investigate the configurations of the Pr form andthe Pr → Pfr conversion mechanisms by theoreticalcalculations. These efforts include the investigation of thestability of different conformations by electronic structurecalculations, the examination of the role of protein environ-ments by molecular dynamics, and the simulation of spectros-copies at different levels of theory.11,12,19−24 These studiesagreed that the Pr → Pfr interconversion in proteinenvironments is triggered by the Z → E photoisomerizationat the C15−C16 bond.In recent years, great efforts have been devoted to clarifying

    the photoinduced isomerization mechanism in phytochromesat the atomic level from chemical aspects. However, such effortsare very challenging because the phytochromes contain verycomplex proteins. Thus, some efforts were devoted to studyingthe chromophores themselves, since they are responsible forthe primary events in the photoinduced biological cycle ofphytochromes.2,14,15,19,23,25−30 In addition, the phytochromechromophores consist of only four open-chain tetrapyrroles,and the nonconjugated side groups should have only minorinfluences on the initial photoabsorption and successiveisomerizations. Thus, some chemists, especially theoreticalscientists, have attempted to choose isolated chromophores astypical models to investigate the photoisomerization mecha-nisms of phytochromes. As the most important photoreceptorin plants, PΦB has often been chosen as such a prototypesystem,2,14,15,19,23,25−30 and its photoisomerization mechanism

    has also been studied without considering the proteinmoiety.2,28,30 Interestingly, these studies found that thephotoconversion mechanism seems to be completely differentwith or without the presence of the original proteinenvironments. Instead of the double-bond twist at the C15−C16 bond, the rotation of ring B and ring C seems to be rathereasier to access in vacuo. Such a difference may be caused bythe steric hindrance of protein environments. As suggested byprevious work,30 ring A and ring B are fixed by the protein-binding pocket, while ring D becomes the “free moiety” thatcan perform the rotation. In vacuo, however, the sterichindrance is wiped off, and all four of the five-memberedrings are free. Furthermore, the electrostatic repulsion betweenrings B and C reinforces their separation.30 Therefore, therotation of ring B or ring C takes place for PΦB.Despite the large number of studies, the primary events in

    the initial step of the photocycle of phytochromes are still notfully understood. Because of the complexity of such systems, itis difficult to examine all of the details from currently availableexperimental and theoretical data. Furthermore, the precursorsproved only that the rotation between ring B and ring C is akey factor for PΦB in the gas phase; they did not determinewhich central bond (C9−C10 or C10−C11) experiences thetwisting motion.1 Some works suggested that the internalrotation around the C10−C11 bond dominates the photo-induced conversion, but their studies did not take the twist ofthe C9−C10 bond into account.28,30 Thus, the elucidation ofthese isomerization processes certainly requires the directexamination of the excited-state dynamics of PΦB.As our first step to investigate the photoinduced reactions of

    phytochromes, we performed nonadiabatic dynamics simu-lations of a PΦB model in vacuo using the on-the-fly surfacehopping method31−33 at the semiempirical OM2/MRCIlevel.34−37 Since a large number of electronic-structurecalculations at a reliable level are required for on-the-flytrajectory calculations, it is important to apply a quantum-chemistry method that can provide a reasonable compromisebetween computational cost and accuracy. For this purpose, theOM2/MRCI method was chosen since many benchmarkcalculations have shown that it can provide a reasonabledescription of molecular excited states with rather lowcomputational cost.34−36 Although it was reported that theOM2/MRCI method may not always be very accurate for somesystems,38 the nonadiabatic dynamics simulation at the OM2/MRCI level should still be a powerful and meaningful tool for

    Figure 1. Chemical structures of biliverdin (BV), phycocyanobilin (PCB), and phytochromobilin (PΦB) in the ZsZsZa configuration.

    The Journal of Physical Chemistry B Article

    dx.doi.org/10.1021/jp408799b | J. Phys. Chem. B 2013, 117, 15976−1598615977

  • many cases after careful benchmark calculations.39−43 Thus,some necessary high-level electronic calculations were alsoperformed for benchmark reasons.In this work, we studied the photoinduced isomerization of a

    PΦB model. For simplicity, only the excited-state dynamicsstarting from the Pr conformation was considered. By trackingthe real-time nuclear motion, we tried to provide a detailedunderstanding of the photoisomerization of the phytochromechromophores in vacuo. This work can become an importantreference for understanding the photoinduced reactions ofphytochrome chromophores within biological environmentsand to determine the roles of the protein residues in the Pr →Pfr photoconversion. In addition, the current work should alsoprovide useful information on the other tetrapyrrole analoguesystems such as porphyrin and its derivatives.

    2. COMPUTATIONAL DETAILS2.1. Molecular Model Construction. Several configu-

    rations exist for PΦB because of the presence of many singleand double C−C bonds. The RR studies of plant phytochromeshave proven that the Pr form of PΦB takes the ZaZsZaconformation.11,12 Thus, the current work also employed theZaZsZa configuration for Pr. To reduce the computational cost,we used the simplified PΦB model reported in refs 28 and 30(Figure 2): some side groups (the C3 thioether linkage, the C8

    and C12 propionic carboxyl groups, and the C2, C7, C13, andC17 methyl groups) were replaced by hydrogen atoms forsimplicity. In principle, these side groups should not beconjugated with the four five-membered rings and shouldhardly influence the conjugation of PΦB. The N atom in ring Cis also protonated as in the earlier reports.28,30 In order toillustrate the photoisomerization procedure conveniently, wedefined several key internal coordinates, such as the importantbond lengths and dihedral angles, as shown in Table 1. Thesymbols d4−5, d5−6, d9−10, d10−11, d14−15, and d15−16 denote thebond lengths of C4−C5, C5−C6, C9−C10, C10−C11, C14−C15, and C15−C16, respectively. The variables τ4−5, τ5−6, τ9−10,τ10−11, τ14−15 and τ15−16 denote the twisting angles around C4−C5, C5−C6, C9−C10, C10−C11, C14−C15, and C15−C16,respectively.2.2. Electronic-Structure Calculations. The semiempir-

    ical calculations were performed using the MNDO99program.44 The Gaussian 0945 and TURBOMOLE 6.546

    programs were also used in the benchmark calculations.2.2.1. Semiempirical Method. The orthogonalization model

    2 (OM2) method was applied to build the semiempiricalelectronic Hamiltonian.34−36 The multireference configurationinteraction (MRCI) method was used to describe the excited-state wave functions.37 The self-consistent field (SCF)

    calculations were performed in the basis of the restrictedopen-shell Hartree-Fock (ROHF) formalism, since it providesthe better treatment of molecular excited states. At the OM2/MRCI level, the potential energies (PEs) of relevant electronicstates and their gradients and nonadiabatic couplings werecalculated analytically. Three references (the closed shell andthe single and double HOMO−LUMO excitations) were usedto build all of the configurations in the MRCI expansion, and allof the single and double excitations from these three referenceswere included in the configuration-interaction treatment. Theactive space (AS) in the MRCI included 16 electrons in 12orbitals [(16,12)]: six π orbitals, two n orbitals and four π*orbitals.Geometry optimizations of the ground-state and the excited-

    state minima were performed using the Newton−Raphsoniteration scheme. The minimum-energy structures of conicalintersections (CI) were obtained using the Lagrange−Newtonmethod.47−49

    The minimum-energy CI geometry (obtained from CIoptimization) was taken as the starting point, and the two-dimensional (2D) rigid scans of PE surfaces were performedalong two relevant dihedral angles. Then the crossing of two PEsurfaces gave us the CI seam within the two-dimensional spacespanned by two dihedral angles. This seam might not be theminimum-energy CI seam within this two-dimensional space,while the former should be a good approximation of the latter.In the discussion below, we used the former choice since itprovides more information on the topology of the two PEsurfaces. As a complementary investigation, we also constructedthe minimum-energy CI seams using constrained CIoptimization by fixing the major reaction coordinates τ9−10and τ10−11, respectively (Figure S-4 in the SupportingInformation).

    2.2.2. High-Level Calculations. In benchmark steps, single-point calculations at critical points were performed at theADC(2)/TZVP level50,51 based on the OM2/MRCI geo-metries. We also attempted to use the DFT/TDDFT method(B3LYP52,53/6-31+G*) to optimize critical points and calculateexcitation energies.

    2.3. Nonadiabatic Dynamics. The photoinduced non-adiabatic decay dynamics was studied by trajectory surface-hopping simulations at the OM2/MRCI level. The non-adiabatic transition was treated using Tully’s fewest-switchesalgorithm.31 A set of initial conditions (geometries andvelocities) was generated from the Wigner distribution functionof normal modes. The initial conditions were then created byputting these snapshots into the S1 state vertically. All of the

    Figure 2. PΦB model in Pr (ZsZsZa) configuration.

    Table 1. Definitions of the Key Geometric Parameters

    definition internal coordinate relative mode

    d4−5 C4−C5 distance C4−C5 bond lengthd5−6 C5−C6 distance C5−C6 bond lengthd9−10 C9−C10 distance C9−C10 bond lengthd10−11 C10−C11 distance C10−C11 bond lengthd14−15 C14−C15 distance C14−C15 bond lengthd15−16 C15−C16 distance C15−C16 bond lengthτ4−5 N23−C4−C5−H5′ dihedral angle twist of C4−C5τ5−6 N22−C6−C5−H5′ dihedral angle twist of C5−C6τ9−10 N22−C9−C10−H10′ dihedral angle twist of C9−C10τ10−11 N21−C11−C10−H10′ dihedral angle twist of C10−C11τ14−15 dihedral angle N21−C14−C15−H15′ twist of C14−C15τ15−16 N20−C16−C15−H15′ dihedral angle twist of C15−C16

    The Journal of Physical Chemistry B Article

    dx.doi.org/10.1021/jp408799b | J. Phys. Chem. B 2013, 117, 15976−1598615978

  • relevant energies, gradients, and nonadiabatic couplings werecalculated analytically in the manner of “on-the-fly”. The steptimes were 0.1 fs for the nuclear motion and 0.001 fs for theelectronic propagation. The final results were obtained byaveraging over all 127 trajectories.

    3. RESULTS AND DISCUSSION

    3.1. Properties of the Pr Conformation. The geometryof the PΦB model in the Pr configuration (ZaZsZa) wasoptimized at the OM2/MRCI level, as shown in Figure S-1 inthe Supporting Information and Table 2. As expected, bothC4−C5 and C15−C16 have double-bond character (∼1.37 Å),while C5−C6 and C14−C15 retain single-bond character(∼1.43 Å). In contrast, the two central bonds (C9−C10 andC10−C11) display similar bond lengths (1.39 and 1.40 Å),reflecting the equal share of the positive charge by rings B andC rather than rings A and D. These geometrical features areconsistent with the data obtained at the B3LYP level andprevious reports.2,28,30 The S0 → S1 vertical energy in the Prconfiguration is 2.28 eV (Table 3), which is also consistent withreference results28,30 and our own benchmark data at theB3LYP/6-31+G* (2.14 eV) and the ADC(2)/TZVP (1.98 eV)levels.Previous references reported that one S1 minimum named

    “PLA” exists in the Franck−Condon region and that itsgeometry is close to the Pr conformation (ZaZsZa). In previouswork, such a minimum was obtained successfully at theTDDFT (TD-B3LYP/SVP, numerical gradient), CIS, andCASSPT2 levels.2,28 Unfortunately, the PLA (S1)min could notbe found at the OM2/MRCI level, and the optimizationstarting from the Pr configuration finally crashed at thegeometry close to the S1/S0 CI characterized by the rotationaround C10−C11 up to ∼90°. To check the discrepancy, weattempted to optimize the PLA at the TDDFT level underdifferent conditions, such as using different density functionals,several types of basis sets, the random phase approximation(RPA) and the Tamm−Dancoff approximation (TDA), and a

    numerical or analytical gradient (Table S-3 in the SupportingInformation). It turned out that the existence of the PLA washighly dependent on above factors. In some cases, we couldoptimize the PLA successfully. For example, the PLA could beoptimized at the PBE0/def2-SVP level using a numerical oranalytical gradient within the RPA and TDA. However, in someother cases (e.g., at the PBE0/TZVP level using a numerical oranalytical gradient within the RPA and TDA), the optimizationdrove the system to leave the Franck−Condon region andperform the C10−C11 twist until the job finally crashed at thegeometry with a very small S1/S0 energy gap. This implies thatthe S1 PE surface in the Franck−Condon region is very flat. Inthis case, even if the PLA exists, this S1 minimum should bevery shallow and not stable. Previous work also found thesimilar feature in their reaction-path scans.2 Since the PEsurface is rather flat in the Franck−Condon region, the excited-state dynamics should not be highly influenced by whether avery shallow S1 minimum exists or not. In view of thereasonable geometries and energies (see section 3.2 below), webelieve that the OM2/MRCI method still can provide a reliabledescription of the molecular motion in the nonadiabaticdynamics of the PΦB model.

    3.2. Potential Energy Surface. Our preliminary non-adiabatic dynamics study found two photoreaction channels:the internal rotation around C9−C10 and the internal rotationaround C10−C11 (see section 3.3). Since the twisting motionof the C15−C16 bond dominates the photochemical reactionof Pr in proteins, it was also considered. The three one-dimensional PE surfaces along the three isomerization paths(C9−C10, C10−C11, and C14−C15) were constructed(Figure S-2 in the Supporting Information). The former twopathways, involving the C9−C10 and C10−C11 twists (FigureS-2a,b) are barrierless from the Pr conformation to thecorresponding CIs, while the third one (Figure S-2c) isunfavored. Previous theoretical studies on the excited-statereaction pathway of the PΦB model in the gas phase alsosuggested a similar mechanism.2,28,30 Thus, in this section our

    Table 2. Internal Coordinates [Bond Distances (Å) and Dihedral Angles (deg)] at Critical Structuresa

    geometry d9−10 d10−11 τ5−6 τ9−10 τ10−11 τ14−15 τ15−16

    Pr 1.39 1.40 −17.9 −167.9 −167.2 −17.5 173.9(S0)α 1.40 1.40 −19.5 1.1 −179.5 17.7 −174.1(S0)β 1.39 1.40 21.3 −177.8 4.0 −19.9 174.1(S1)α 1.45 1.38 −8.5 −90.9 −177.6 14.0 −173.5(S1)β 1.38 1.45 −18.1 −176.7 −92.0 −7.4 170.4CI01α‑min 1.45 1.38 −7.8 −91.5 −177.2 −13.1 173.7CI01β‑min 1.38 1.45 −21.1 −176.1 −93.1 −6.0 −167.1

    aThe structure of (S0)α was optimized from the geometry obtained by twisting the C9−C10 bond about 180° from the Pr conformation; (S0)β wasobtained by performing the same procedure on the C10−C11 bond.

    Table 3. S0 and S1 Energies (eV), S0 → S1 Vertical Excitation Energies (eV), and Oscillator Strengths (in Parentheses) forCritical Geometries at the OM2/MRCI, B3LYP/6-31+G*, and ADC(2)/TZVP Levels

    OM2/MRCI B3LYP/6-31+G* ADC(2)/TZVP

    geometry S0 S1 S0 → S1 S0 S1 S0 → S1 S0 S1 S0 → S1

    Pr 0.10 2.38 2.28 (1.04) 0.07 2.21 2.14 (1.70) 0.03 1.98 1.95 (1.56)(S0)α 0.00 2.30 2.30 (1.07) 0.01 2.14 2.13 (1.98) 0.01 1.96 1.95 (1.84)(S0)β 0.08 2.44 2.36 (1.10) 0.00 2.13 2.13 (1.91) 0.00 1.97 1.97 (1.83)(S1)α 1.43 1.56 0.13 (0.00) 1.43 1.82 0.39 (0.00)(S1)β 1.31 1.55 0.24 (0.00) 1.28 1.86 0.58 (0.00)CI01α‑min 1.58CI01β‑min 1.56

    The Journal of Physical Chemistry B Article

    dx.doi.org/10.1021/jp408799b | J. Phys. Chem. B 2013, 117, 15976−1598615979

  • work focused only on the PE surfaces along reactivecoordinates responsible for two observed major channels(internal rotations around C9−C10 and C10−C11) and didnot address others.3.2.1. Channel I. One S0 minimum with the ZsZsZa

    conformation, (S0)α, is located at the OM2/MRCI level whenthe whole A−B double-ring moiety rotates around the C9−C10bond by about 170° (Figure 3 and Table 2). The energy of(S0)α is 0.10 eV lower than that of the Pr conformation at theOM2/MRCI level, and the S0 → S1 vertical excitation energy is2.30 eV at (S0)α. Since similar data were also obtained at theADC(2)/TZVP and B3LYP/6-31+G* levels (Table 3), theOM2/MRCI result is reliable.At the OM2/MRCI level, one S0/S1 CI minimum (CI01α‑min,

    1.68 eV above Pr) is located by the rotation of the A−B double-ring moiety around the C9−C10 bond by about 80° from thePr conformation (Figure 3 and Tables 2 and 3). From the Pr toCI01α‑min, the other central C−C bond (C10−C11) performs

    only a weak twisting (τ10−11 from −167.2° to −177.2°). Duringthis rotation, the angle between rings A and B becomes smaller(τ5−6 from −17.9° to −7.8°). This indicates that both thetwisting motion of the central C10−C11 bond and thereduction of the ring A/ring B angles are required to accessCI01α‑min. The C9−C10 bond distance increases from 1.39 to1.45 Å, while the C10−C11 bond distance becomes shorterfrom 1.40 to 1.38 Å. Therefore, C9−C10 and C10−C11 tendto become single and double bonds, respectively. At thisgeometry, the intramolecular charge transfer is indicated by themolecular orbitals (Figure S-3 in the Supporting Information)and net atomic charge distributions (Table S-5 in theSupporting Information), also consistent with the referencereports.28,30

    One S1 minimum, (S1)α, is present in the vicinity of CI01α‑min.The major geometry distinction between (S1)α and CI01α‑min isthat they show different angles between rings C and D. FromPr to (S1)α, the twisting angle of the C14−C15 bond (τ9−10)

    Figure 3. Optimized geometries of two ground-state minima [(S0)α and (S0)β], two lowest singlet excited-state minima [(S1)α and (S1)β], and twoS1/S0 conical intersection minima [CI01α‑min and CI01β‑min] at the OM2/MRCI level of theory.

    Figure 4. Reaction paths from the Pr conformation to (S0)α. (a) Two-dimensional PE surfaces as functions of τ9−10 and τ5−6. (b) Profile of the CI01αseam in the space spanned by τ9−10 and τ5−6. (c, d) PE surfaces of the CI01α seam as functions of τ9−10 and τ5−6, respectively. τ5−6 and τ9−10 in the plotdenote the dihedral angles around C5−C6 and C9−C10, respectively. “+” in the plot refers to the location of CI01α‑min, and “*” in the plot shows thevertical excitation energy of Pr.

    The Journal of Physical Chemistry B Article

    dx.doi.org/10.1021/jp408799b | J. Phys. Chem. B 2013, 117, 15976−1598615980

  • changes from −17.5° to 14.0°, which is much bigger than thechange from Pr to CI01α‑min (τ9−10 from −17.5° to −13.1°). Atthe (S1)α geometry, the S0−S1 energy gap is very small (only∼0.13 eV), which indicates that (S1)α is close to the S0/S1 CI.Since DFT failed at the configuration with a small S0−S1 energygap, all attempts to optimize (S1)α at the TDDFT level crashedno matter which functional was selected. We performed onlysingle-point ADC(2) calculations at the (S1)α geometry in thebenchmark step. At the ADC(2) level of theory, the S0−S1energy gap was also found to become very small (Table 3)It is well-known that nonadiabatic transitions do not

    necessarily take place at the minimum-energy geometry of aCI. Instead, all accessible regions of the CI seam should beresponsible. Thus, we attempted to check the topology of theCI seam. The 2D PE surfaces were constructed by performingrigid scans along internal coordinates τ9−10 and τ5−6 startingfrom the Pr form (Figure 4).Starting from the Pr conformation, it is easy to access CI01α

    in the S1 state because this channel is almost barrierless (Figure4a). It should be possible to derive the reaction mechanism ofchannel I from the PE surface profiles and the CI topology.After the molecule is excited to the S1 state at the Prconformation, it relaxes to the CI01α seam by rotation of ring Baround the C9−C10 bond (τ9−10) and ring A around the C5−C6 bond (τ5−6). After its decay to the S0 state at the CI01α seam,the molecule may experience further rotation of ring B forwardto (S0)α or backward to Pr ultimately. Interestingly, the CI01αseam exhibits a line in the two-dimensional space spanned byτ9−10 and τ5−6 (Figure 4b). It is also interesting to notice thatthe vertical excitation energy of Pr is above a part of the CIseam (range from 2.28 to 1.58 eV in Figure 4c,d). Thus, theexcited-state PE surface may drive the system toward not onlyCI01α‑min but also the CI01α seam. In other words, the anglebetween rings B and A may also play an essential role in the

    excited-state dynamics in addition to the twisting motion of theC9−C10 bond.

    3.2.2. Channel II. At the OM2/MRCI level, one S0minimum displaying the ZaEsZs conformation, (S0)β, is alsoobtained by rotation of the whole C−D double-ring moiety byabout 170° around the C10−C11 bond (Figure 3 and Table 2).All three S0 minima in the current work [Pr, (S0)α, and (S0)β]are located in a similar energy range at the OM2/MRCI,B3LYP/6-31+G*, and ADC(2)/TZVP levels. For (S0)β, thevertical excitation energy of S1 is 2.36 eV at the OM2/MRCIlevel, consistent with the B3LYP/6-31+G* (2.13 eV) andADC(2)/TZVP (1.97 eV) results. Similar to channel I, theconsistent data obtained at the OM2/MRCI, DFT/TDDFT,and ADC(2)/TZVP levels proves that the semiempirical OM2/MRCI method gives a reasonable description of the excitedstates for the PΦB model.At the OM2/MRCI level, another S0/S1 CI minimum was

    located (CI01β‑min, 1.66 eV above Pr), which is responsible forthe nonadiabatic decay via channel II. CI01β‑min is characterizedby rotation of the whole C−D double-ring moiety around theC10−C11 bond by about 80° from the Pr conformation(Figure 3 and Tables 2 and 3). During this rotation, the anglebetween rings C and D also decreases by around 10° (τ14−15from −17.5° to −6.0°). In this situation, the C10−C11 bondlength increases to 1.45 Å, while the C9−C10 bond becomesshorter to 1.38 Å. Compared with Pr, C10−C11 and C9−C10tend to become single and double bonds, respectively, atCI01α‑min. The intramolecular charge transfers also occur at thisgeometry. The molecular orbitals (Figure S-3 in the SupportingInformation) and net atomic charge distributions (Table S-5 inthe Supporting Information) support such a conclusion.One S1 minimum, (S1)β, also exists in the vicinity of CI01β‑min.

    The major geometry distinction of (S1)β and CI01β‑min is thedifferent C15−C16 twisting angles (τ15−16). From Pr to (S1)β,τ15−16 changes from 173.9° to 170.4°, which is much less than

    Figure 5. Reaction paths from the Pr conformation to (S0)β. (a) Two-dimensional PE surfaces as functions of τ10−11 and τ14−15. (b) Profile of theCI01β seam in the space spanned by τ10−11 and τ14−15. (c, d) PE surfaces of the CI01β seam as functions of τ10−11 and τ14−15, respectively. τ10−11 andτ14−15 in the plot denote the dihedral angles around C10−C11 and C14−C15, respectively. “+” in the plot refers to the location of CI01β‑min, and “*”in the plot shows the vertical excitation energy of Pr.

    The Journal of Physical Chemistry B Article

    dx.doi.org/10.1021/jp408799b | J. Phys. Chem. B 2013, 117, 15976−1598615981

  • the change from Pr to CI01β‑min (173.9° to −167.1°). Similar to(S1)α, the S0−S1 energy gap of the two states is also very small(only ∼0.24 eV) at the (S1)β geometry, which is also close tothe S0/S1 CI as discussed above (Table 3).For illustration, we also constructed the 2D PE surfaces by

    performing rigid scans along the internal coordinates τ10−11 andτ14−15 starting from the Pr form (Figure 5). Because the basicfeatures of the two-dimensional S0 and S1 surfaces for channelII are very similar to those of channel I, a similar reactionmechanism should be expected. From the Pr conformation, thesystem can quickly move to the CI01β seam, instead of onlyCI01β‑min (Figure 5c,d). After the internal conversion to the S0state, the system may move forward to (S0)β or backward to thePr conformer. Similar to channel I, both the C10−C11 (τ10−11)and C14−C15 (τ14−15) bonds may perform the twisting motionin the excited-state dynamics.Generally speaking, both channels may be easily accessible

    because of the lack of an obvious barrier, while the finaldestinations of the nonadiabatic dynamics may be quitedifferent. When the molecule decays to the S0 state via theCI seams, the molecule may rotate forward to the other S0minimum [(S0)α or (S0)β] or just rotate backward to theoriginal Pr conformation. In order to get further insight into thereal-time molecular motions for the two channels, the ultrafastnonadiabatic dynamics study is necessary.3.3. Nonadiabatic Dynamics Study of PΦB. A set of

    trajectories was used in the surface-hopping nonadiabaticdynamics study at the OM2/MRCI level. Starting from thePr configuration, the fractional occupation of the S1 state showsan ultrafast decay (Figure 6). More than 50% of the trajectoriesjump to the S0 state within 300 fs, and most of the trajectoriesdecay to the S0 state within 500 fs.

    To inspect the details of the nonadiabatic transitions, thegeometrical distributions of two key internal coordinates (τ9−10and τ10−11) were constructed for all of the initial geometries andhopping geometries, as shown in Figure 7. Because all of theinitial geometries are sampled in the vicinity of the Prconfiguration, the absolute values of the angles around bothcentral C−C bonds (C9−C10 and C10−C11) between rings Band C are distributed around 170° at t = 0 (Figure 7a). Thentrajectories move on the S1 state and access the CI seams (CI01αor CI01β), where the S1 → S0 hop takes place. Clearly, tworeaction channels can be identified immediately in Figure 7b.Many trajectories (74.1%) pass CI01α by a twisting motion ofthe C9−C10 bond, while some trajectories (22.8%) accessCI01β with the twisting motion of the C10−C11 bond. We didnot observe other motions such as the Hula twist of the C9−C10 and C10−C11 bonds.

    As shown in the reaction pathways, the trajectories via thetwo CIs result in quite different products (Figure 8). Many

    trajectories (74.1%) jump back to the ground state via CI01α.After the S1 → S0 hop, two-thirds of these trajectories returnback to the Pr conformation (46.5%) and only one-third(27.6%) move toward to (S0)α. There are also many trajectories(22.8%) passing CI01β. Among them, the numbers oftrajectories that move backward to Pr (11.0%) and forwardto (S0)β (11.8%) are similar. Only a very small number oftrajectories do not decay to the S0 state (3.1%), and thus, thischannel is negligible. For the current system, we did notobserve any trajectory passing other reaction channelscharacterized by the significant C4−C5 or C15−C16 twist.This finding is consistent with the mechanism suggested byprevious reaction-path calculations at various levels ofelectronic-structure theory.2,28,30

    Since the CI seams (not only the CI minima) are accessiblein the excited state, it is necessary to determine the relevant

    Figure 6. Time-dependent average fractional occupations of electronicstates for nonadiabatic dynamics initiated from the Pr conformation.

    Figure 7. Geometrical distributions as functions of the C9−C10(τ9−10) and C10−C11 (τ10−11) angles over (a) all initial geometriesand (b) hopping geometries.

    Figure 8. Ratio distributions of trajectories toward different reactionchannels. “α+ra” denotes the trajectories passing CI01α and moving to(S0)α; “α+Pr” refers to the trajectories passing CI01α and rotating backto Pr conformation; “β+rb” denotes the trajectories passing CI01β andmoving to (S0)β; “β+Pr” refers to trajectories passing CI01β androtating back to Pr conformation; “no hop” refers to the trajectoriesthat did not hop.

    The Journal of Physical Chemistry B Article

    dx.doi.org/10.1021/jp408799b | J. Phys. Chem. B 2013, 117, 15976−1598615982

  • internal coordinates involved in the nonadiabatic dynamics. Forthis purpose, we investigated the details of the nucleardistributions at different time.For channel I, we plotted the time-dependent geometrical

    distributions as functions of the internal coordinates τ9−10 andτ5−6, which describe the rotations around C9−C10 and C5−C6respectively (Figure 9). Within the first 100 fs, the trajectories

    remain the Pr conformation. At 300 fs, many trajectories reachthe CI01α seam region. During this process, the twistingmotions of the C9−C10 and C5−C6 bonds are observed. Suchdetailed molecular motion can be examined again carefullyfrom Figure 10. At the Pr configuration, the twisting angles ofC9−C10 (τ9−10, between rings B and C) and C5−C6 (τ5−6,between rings A and B) are −167.9° and −17.9°, respectively.At a typical hopping geometry, these angles become −114.8°and −13.8°, respectively. Thus, rings A and B cannot be treatedas a single moiety performing an overall rotation during theexcited-state motion; otherwise, twisting motion of C5−C6should not be observed. Instead, the rotations of rings A and Btake place in a nonsynchronous way. At CI01α‑min, the above twoangles become −91.5° and −7.8°, respectively. The tworelevant angles (the C9−C10 twisting angle and the ring A/ring B angle) show quite a difference at the above typicalhopping geometry and CI01α‑min. Thus, not the CI minimumgeometry (CI01α‑min) but the CI01α seam (Figures 4 and 10) isresponsible for the nonadiabatic decay of the PΦB model. Afterthe S1 → S0 hop, the forward and backward internal rotationsaround the C9−C10 bond finally result in the photoproduct(S0)α and the original Pr conformation, respectively.For channel II, we plotted the time-dependent geometrical

    distributions of the relevant internal coordinates τ10−11 andτ14−15, which refer to the rotations around the C10−C11 and

    C14−C15 bonds, respectively, as shown in Figure 11. Similar tochannel I, in the early stage of the dynamics, all of thetrajectories remain the Pr conformation. As time progresses, theinternal rotations around C10−C11 (τ10−11) and C14−C15(τ14−15) are involved in the motion (Figure 12). Similarly, theCI seam (instead of only CI01β‑min) is reached. After the S1→S0hop, the trajectories also show two different types of motions.The backward internal rotation around C10−C11 gives the Prconformation, while the successive forward internal rotationaround C10−C11 results in the (S0)β product.In our nonadiabatic dynamics study, the model compound of

    the PΦB in the gas phase displays a different reactionmechanism from the actual chromophore in protein environ-

    Figure 9. Geometrical distributions as functions of the twisting anglesaround C9−C10 (τ9−10) and C5−C6 (τ5−6) at different time steps forall of the trajectories passing CI01α. The white lines in the plots refer tothe positions of the CI01α seam.

    Figure 10. Structures of Pr, CI01α‑min, and a typical hopping geometry.

    Figure 11. Geometrical distributions as functions of the twistingangles around C10−C11 (τ10−11) and C14−C15 (τ14−15) at differenttime steps for all of the trajectories passing CI01β. The white lines inthe plots refer to the positions of the CI01β seam.

    The Journal of Physical Chemistry B Article

    dx.doi.org/10.1021/jp408799b | J. Phys. Chem. B 2013, 117, 15976−1598615983

  • ments. Consistent with the references’ reports,2,28,30 we did notobserve the pathways along the rotation around the C4−C5 orC15−C16 bonds. Instead, only two reaction channels werefound: the rotations around C9−C10 and C5−C6 or therotations around C10−C11 and C14−C15. Both of thesereaction channels seem to be easily accessed because of the lackof an obvious barrier. For the PΦB model, the C−D double-ring moiety is much heavier than the A−B double-ring moietybecause of the ethenyl group in ring D, so the former has thelarger moment of inertia. This may explain why channel I ispreferred over channel II.Obviously, the one-dimensional picture used to understand

    the reaction mechanism (including the CI minimum, the one-dimensional reaction pathway, and leading reaction coordi-nates) is still qualitatively reasonable. This idea can befurthermore supported by the reaction pathway (Figure S-2in the Supporting Information) and hopping geometrydistribution (Figure S-4 in the Supporting Information).However, a more precise mechanism would involve moredetails, such as the involvement of other nuclear motions andthe topology of the CI seams. As discussed in the literature, thisfeature seems to widely exist in different molecular systems.54

    The deep understanding of this issue certainly requires anexploration of the topologies of complex multidimensional PEsurfaces and subsequent nonadiabatic dynamics.54

    4. CONCLUSIONS

    The primary event of the photoinduced isomerization reactionof the PΦB model in vacuo was simulated using surface-hopping nonadiabatic dynamics at the semiempirical OM2/MRCI level. Two reaction channels (via CI01α and CI01β) werefound to be responsible for the excited-state decay, which aremainly characterized by the rotations around the C9−C10 andC10−C11 bonds, respectively.In the surface-hopping simulations, the decay of the S1

    population is an ultrafast process (within 500 fs) because ofthe lack of barriers on the reaction pathways. The reactionmechanism is described as shown in Figure 13. Mosttrajectories pass CI01α (74.1%), which is characterized by thetwisting motion of the C9−C10 bond, while less than 25% ofthe trajectories pass CI01β, which displays the twisting motionof the C10−C11 bond. For the CI01α channel, after thetrajectories decay to the S0 state, only one-third of them moveforward to (S0)α while two-thirds of them move backward toPr. For the CI01β channel, after the trajectories decay to the S0state, the numbers of trajectories following the forward channel[to (S0)β] and the backward channel (to Pr) seem rathersimilar. No matter which pathway is followed, the nonadiabaticdecay always involves a part of the CI seam instead of only theCI minimum. Therefore, the molecular motions on the excitedstate also involve other twisting motions, for instance therotations of rings A and D. In this sense, we can conclude thatthe simple minimum-energy reaction pathway calculation is notenough to correctly locate the internal coordinates responsiblefor the nonadiabatic decay for PΦB in the gas phase.Although our work can provide a very clear picture of the

    photoisomerization process of PΦB in vacuo, we know that thereaction dynamics is different from that in protein environ-ments. This means that the reaction mechanism may bemodified by the special constraints of protein environments,including steric hindrance, hydrogen-bonding networks, aniso-tropic electrostatic interactions, and charge status. In suchprotein environments, the steric interactions may block the fastinternal conversion processes that exist in the gas phase. Thus,to simulate the photoinduced Pr → Pfr conversion of PΦB inplants, the protein environments must be considered. In thiscase, the hybrid QM/MM approach may be a suitable choice todeal with such complex reactions. This certainly represents avery challenging task and will be the subject of future research.

    Figure 12. Structures of Pr, CI01β‑min, and a typical hopping geometry.

    Figure 13. Final mechanistic scheme and quantum yields of photochemical reactions of PΦB from the Pr conformation in the gas phase.

    The Journal of Physical Chemistry B Article

    dx.doi.org/10.1021/jp408799b | J. Phys. Chem. B 2013, 117, 15976−1598615984

  • ■ ASSOCIATED CONTENT*S Supporting InformationFigures showing the Pr geometry, one-dimensional potentialsurfaces, molecular orbitals, and the distribution of hoppinggeometries and tables of representative internal coordinates,optimization results at the PLA at TDDFT levels, geometrydistributions for different reaction channels, net atomic charges,and Cartesian coordinates of all important geometries. Thismaterial is available free of charge via the Internet at http://pubs.acs.org.

    ■ AUTHOR INFORMATIONCorresponding Author*Fax: +86-532-80662778. Tel: +86-532-80662630. E-mail:[email protected] authors declare no competing financial interest.

    ■ ACKNOWLEDGMENTSThis work was supported by the CAS 100 Talents Project, theNational Natural Science Foundation of China (Grant Nos.21103213 and 91233106), and the Director InnovationFoundation of CAS-QIBEBT. The authors thank the Super-computing Center, Computer Network Information Center,CAS, and the supercomputational center of CAS-QIBEBT forproviding computational resources.

    ■ REFERENCES(1) Rockwell, N. C.; Su, Y.-S.; Lagarias, J. C. Phytochrome Structureand Signaling Mechanisms. Annu. Rev. Plant Biol. 2006, 57, 837−858.(2) Altoe,̀ P.; Climent, T.; De Fusco, G. C.; Stenta, M.; Bottoni, A.;Serrano-Andrieś, L.; Merchań, M.; Orlandi, G.; Garavelli, M.Deciphering Intrinsic Deactivation/Isomerization Routes in aPhytochrome Chromophore Model. J. Phys. Chem. B 2009, 113,15067−15073.(3) Alvey, R. M.; Biswas, A.; Schluchter, W. M.; Bryant, D. A.Attachment of Noncognate Chromophores to CpcA of Synechocystissp. PCC 6803 and Synechococcus sp. PCC 7002 by HeterologousExpression in Escherichia coli. Biochemistry 2011, 50, 4890−4902.(4) Bongards, C.; Gar̈tner, W. The Role of the Chromophore in theBiological Photoreceptor Phytochrome: An Approach Using Chemi-cally Synthesized Tetrapyrroles. Acc. Chem. Res. 2010, 43, 485−495.(5) Wagner, J. R.; Brunzelle, J. S.; Forest, K. T.; Vierstra, R. D. ALight-Sensing Knot Revealed by the Structure of the Chromophore-Binding Domain of Phytochrome. Nature 2005, 438, 325−331.(6) Wagner, J. R.; Zhang, J.; Brunzelle, J. S.; Vierstra, R. D.; Forest, K.T. High Resolution Structure of Deinococcus BacteriophytochromeYields New Insights into Phytochrome Architecture and Evolution. J.Biol. Chem. 2007, 282, 12298−12309.(7) Yang, X.; Stojkovic,́ E. A.; Kuk, J.; Moffat, K. Crystal Structure ofthe Chromophore Binding Domain of an Unusual Bacteriophyto-chrome, RpBphP3, Reveals Residues That Modulate Photoconversion.Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 12571−12576.(8) Essen, L.-O.; Mailliet, J.; Hughes, J. The Structure of a CompletePhytochrome Sensory Module in the Pr Ground State. Proc. Natl.Acad. Sci. U.S.A. 2008, 105, 14709−14714.(9) Cornilescu, G.; Ulijasz, A. T.; Cornilescu, C. C.; Markley, J. L.;Vierstra, R. D. Solution Structure of a Cyanobacterial PhytochromeGAF Domain in the Red-Light-Absorbing Ground State. J. Mol. Biol.2008, 383, 403−413.(10) Van Thor, J. J.; Mackeen, M.; Kuprov, I.; Dwek, R. A.; Wormald,M. R. Chromophore Structure in the Photocycle of the CyanobacterialPhytochrome Cph1. Biophys. J. 2006, 91, 1811−1822.(11) Mroginski, M. A.; Murgida, D. H.; von Stetten, D.; Kneip, C.;Mark, F.; Hildebrandt, P. Determination of the Chromophore

    Structures in the Photoinduced Reaction Cycle of Phytochrome. J.Am. Chem. Soc. 2004, 126, 16734−16735.(12) Mroginski, M. A.; Murgida, D. H.; Hildebrandt, P. TheChromophore Structural Changes During the Photocycle ofPhytochrome: A Combined Resonance Raman and QuantumChemical Approach. Acc. Chem. Res. 2007, 40, 258−266.(13) Ulijasz, A. T.; Cornilescu, G.; Cornilescu, C. C.; Zhang, J.;Rivera, M.; Markley, J. L.; Vierstra, R. D. Structural Basis for thePhotoconversion of a Phytochrome to the Activated Pfr Form. Nature2010, 463, 250−254.(14) Andel, F.; Lagarias, J. C.; Mathies, R. A. Resonance RamanAnalysis of Chromophore Structure in the Lumi-R Photoproduct ofPhytochrome. Biochemistry 1996, 35, 15997−16008.(15) Durbeej, B.; Borg, O. A.; Eriksson, L. A. Phytochromobilin C15-Z,syn → C15-E,anti Isomerization: Concerted or Stepwise? Phys.Chem. Chem. Phys. 2004, 6, 5066−5073.(16) Fischer, A. J.; Lagarias, J. C. Harnessing Phytochrome’s GlowingPotential. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 17334−17339.(17) Freer, L. H.; Kim, P. W.; Corley, S. C.; Rockwell, N. C.; Zhao,L.; Thibert, A. J.; Lagarias, J. C.; Larsen, D. S. ChemicalInhomogeneity in the Ultrafast Dynamics of the DXCF Cyanobacter-iochrome Tlr0924. J. Phys. Chem. B 2012, 116, 10571−10581.(18) Hahn, J.; Strauss, H. M.; Schmieder, P. Heteronuclear NMRInvestigation on the Structure and Dynamics of the ChromophoreBinding Pocket of the Cyanobacterial Phytochrome Cph1. J. Am.Chem. Soc. 2008, 130, 11170−11178.(19) Hasegawa, J.-y.; Isshiki, M.; Fujimoto, K.; Nakatsuji, H.Structure of Phytochromobilin in the Pr and Pfr Forms: SAC-CITheoretical Study. Chem. Phys. Lett. 2005, 410, 90−93.(20) Murgida, D. H.; von Stetten, D.; Hildebrandt, P.; Schwinte,́ P.;Siebert, F.; Sharda, S.; Ga ̈rtner, W.; Mroginski, M. A. TheChromophore Structures of the Pr States in Plant and BacterialPhytochromes. Biophys. J. 2007, 93, 2410−2417.(21) Matute, R. A.; Contreras, R.; Peŕez-Hernańdez, G.; Gonzaĺez, L.The Chromophore Structure of the Cyanobacterial PhytochromeCph1 As Predicted by Time-Dependent Density Functional Theory. J.Phys. Chem. B 2008, 112, 16253−16256.(22) Kaminski, S.; Mroginski, M. A. Molecular Dynamics ofPhycocyanobilin Binding Bacteriophytochromes: A Detailed Study ofStructural and Dynamic Properties. J. Phys. Chem. B 2010, 114,16677−16686.(23) Matute, R. A.; Contreras, R.; Gonzaĺez, L. Time-DependentDFT on Phytochrome Chromophores: A Way to the RightConformer. J. Phys. Chem. Lett. 2010, 1, 796−801.(24) Mroginski, M. A.; Kaminski, S.; von Stetten, D.; Ringsdorf, S.;Ga ̈rtner, W.; Essen, L.-O.; Hildebrandt, P. Structure of theChromophore Binding Pocket in the Pr State of Plant PhytochromephyA. J. Phys. Chem. B 2011, 115, 1220−1231.(25) Borg, O. A.; Durbeej, B. Which Factors Determine the Acidityof the Phytochromobilin Chromophore of Plant Phytochrome? Phys.Chem. Chem. Phys. 2008, 10, 2528−2537.(26) Borg, O. A.; Durbeej, B. Relative Ground and Excited-State pKaValues of Phytochromobilin in the Photoactivation of Phytochrome: AComputational Study. J. Phys. Chem. B 2007, 111, 11554−11565.(27) Durbeej, B.; Eriksson, L. A. Protein-Bound ChromophoresAstaxanthin and Phytochromobilin: Excited State Quantum ChemicalStudies. Phys. Chem. Chem. Phys. 2006, 8, 4053−4071.(28) Durbeej, B. On the Primary Event of Phytochrome: QuantumChemical Comparison of Photoreactions at C4, C10 and C15. Phys.Chem. Chem. Phys. 2009, 11, 1354−1361.(29) Durbeej, B.; Borg, O. A.; Eriksson, L. A. ComputationalEvidence in Favor of a Protonated Chromophore in the Photo-activation of Phytochrome. Chem. Phys. Lett. 2005, 416, 83−88.(30) Strambi, A.; Durbeej, B. Initial Excited-State Relaxation of theBilin Chromophores of Phytochromes: A Computational Study.Photochem. Photobiol. Sci. 2011, 10, 569−579.(31) Tully, J. C. Molecular Dynamics with Electronic Transitions. J.Chem. Phys. 1990, 93, 1061−1071.

    The Journal of Physical Chemistry B Article

    dx.doi.org/10.1021/jp408799b | J. Phys. Chem. B 2013, 117, 15976−1598615985

    http://pubs.acs.orghttp://pubs.acs.orgmailto:[email protected]

  • (32) Hammes-Schiffer, S.; Tully, J. C. Proton Transfer in Solution:Molecular Dynamics with Quantum Transitions. J. Chem. Phys. 1994,101, 4657−4667.(33) Fabiano, E.; Keal, T. W.; Thiel, W. Implementation of SurfaceHopping Molecular Dynamics using Semiempirical Methods. Chem.Phys. 2008, 349, 334−347.(34) Weber, W.; Thiel, W. Orthogonalization Corrections forSemiempirical Methods. Theor. Chem. Acc. 2000, 103, 495−506.(35) Weber, W. Ein Neues Semiempirisches NDDO-Verfahren mitOrthogonalisierungskorrekturen: Entwicklung des Modells, Imple-mentierung, Parametrisierung und Anwendungen, 1996.(36) Otte, N.; Scholten, M.; Thiel, W. Looking at Self-Consistent-Charge Density Functional Tight Binding from a SemiempiricalPerspective. J. Phys. Chem. A 2007, 111, 5751−5755.(37) Koslowski, A.; Beck, M. E.; Thiel, W. Implementation of aGeneral Multireference Configuration Interaction Procedure withAnalytic Gradients in a Semiempirical Context Using the GraphicalUnitary Group Approach. J. Comput. Chem. 2003, 24, 714−726.(38) Barbatti, M.; Lan, Z.; Crespo-Otero, R.; Szymczak, J. J.; Lischka,H.; Thiel, W. Critical Appraisal of Excited State NonadiabaticDynamics Simulations of 9H-adenine. J. Chem. Phys. 2012, 137,No. 22A503.(39) Kazaryan, A.; Lan, Z.; Schaf̈er, L. V.; Thiel, W.; Filatov, M.Surface Hopping Excited-State Dynamics Study of the Photo-isomerization of a Light-Driven Fluorene Molecular Rotary Motor. J.Chem. Theory Comput. 2011, 7, 2189−2199.(40) Zhuang, X.; Wang, J.; Lan, Z. Photoinduced NonadiabaticDecay and Dissociation Dynamics of Dimethylnitramine. J. Phys.Chem. A 2013, 117, 4785−4793.(41) Lan, Z.; Fabiano, E.; Thiel, W. Photoinduced NonadiabaticDynamics of Pyrimidine Nucleobases: On-the-Fly Surface-HoppingStudy with Semiempirical Methods. J. Phys. Chem. B 2009, 113, 3548−3555.(42) Lan, Z.; Lu, Y.; Weingart, O.; Thiel, W. Nonadiabatic DecayDynamics of a Benzylidene Malononitrile. J. Phys. Chem. A 2012, 116,1510−1518.(43) Zhang, W.; Lan, Z.; Sun, Z.; Gaffney, K. J. Resolving Photo-Induced Twisted Intramolecular Charge Transfer with VibrationalAnisotropy and TDDFT. J. Phys. Chem. B 2012, 116, 11527−11536.(44) Thiel, W. MNDO Program, version 6.1, 2007.(45) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Menucci, B.;Petersson, G. A.; Nakatsuji, H.; Caricato, M.; et al. Gaussian 09;Gaussian, Inc.: Wallingford, CT, 2009.(46) Ahlrichs, R.; Bar̈, M.; Has̈er, M.; Horn, H.; Kölmel, C.Electronic Structure Calculations on Workstation Computers: TheProgram System Turbomole. Chem. Phys. Lett. 1989, 162, 165−169.(47) Bearpark, M. J.; Robb, M. A.; Schlegel, H. B. A Direct Methodfor the Location of the Lowest Energy Point on a Potential SurfaceCrossing. Chem. Phys. Lett. 1994, 223, 269−274.(48) Yarkony, D. R. Diabolical Conical Intersections. Rev. Mod. Phys.1996, 68, 985−1013.(49) Keal, T.; Koslowski, A.; Thiel, W. Comparison of Algorithms forConical Intersection Optimisation Using Semiempirical Methods.Theor. Chem. Acc. 2007, 118, 837−844.(50) Schirmer, J. Beyond the Random-Phase Approximation: A NewApproximation Scheme for the Polarization Propagator. Phys. Rev. A1982, 26, 2395−2416.(51) Schafer, A.; Huber, C.; Ahlrichs, R. Fully Optimized ContractedGaussian Basis Sets of Triple Zeta Valence Quality for Atoms Li to Kr.J. Chem. Phys. 1994, 100, 5829−5835.(52) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle−Salvetti Correlation-Energy Formula into a Functional of the ElectronDensity. Phys. Rev. B 1988, 37, 785−789.(53) Becke, A. D. Density-Functional Thermochemistry. III. TheRole of Exact Exchange. J. Chem. Phys. 1993, 98, 5648−5652.(54) Robb, M. A. Conical Intersections in Organic Photochemistry.In Conical Intersections II: Theory, Computation and Experiment;

    Domcke, W., Yarkony, D. R., Köppel, H., Eds.; World Scientific:Singapore, 2011; Chapter 1.

    The Journal of Physical Chemistry B Article

    dx.doi.org/10.1021/jp408799b | J. Phys. Chem. B 2013, 117, 15976−1598615986