Background: Bacteriorhodopsin (BR) functions as a proton pump, whereas Anabaena sensory rhodopsin (ASR) functions as a photosensor.
Results: pKa of the conserved Asp residues near the Schiff base significantly differ between BR and ASR.
Conclusion: The O-D stretching frequencies for D2O are correlated with pKa(Asp).
Significance: The presence of a strongly H-bonded water results from the proton-pumping activity in BR.
Keywords: Crystal Structure, Fourier Transform IR (FTIR), Proton Pumps, Quantum Chemistry, Rhodopsin
Abstract
Bacteriorhodopsin (BR) functions as a light-driven proton pump, whereas Anabaena sensory rhodopsin (ASR) is believed to function as a photosensor despite the high similarity in their protein sequences. In Fourier transform infrared (FTIR) spectroscopic studies, the lowest O-D stretch for D2O was observed at ∼2200 cm−1 in BR but was significantly higher in ASR (>2500 cm−1), which was previously attributed to a water molecule near the Schiff base (W402) that is H-bonded to Asp-85 in BR and Asp-75 in ASR. We investigated the factors that differentiate the lowest O-D stretches of W402 in BR and ASR. Quantum mechanical/molecular mechanical calculations reproduced the H-bond geometries of the crystal structures, and the calculated O-D stretching frequencies were corroborated by the FTIR band assignments. The potential energy profiles indicate that the smaller O-D stretching frequency in BR originates from the significantly higher pKa(Asp-85) in BR relative to the pKa(Asp-75) in ASR, which were calculated to be 1.5 and −5.1, respectively. The difference is mostly due to the influences of Ala-53, Arg-82, Glu-194–Glu-204, and Asp-212 on pKa(Asp-85) in BR and the corresponding residues Ser-47, Arg-72, Ser-188-Asp-198, and Pro-206 on pKa(Asp-75) in ASR. Because these residues participate in proton transfer pathways in BR but not in ASR, the presence of a strongly H-bonded water molecule near the Schiff base ultimately results from the proton-pumping activity in BR.
Introduction
Bacteriorhodopsin (BR)2 functions as a light-driven proton pump, and the driving force for its action is provided by photoisomerization of the all-trans retinal chromophore, which is covalently attached to Lys-216 via the Schiff base, to 13-cis. This leads to proton transfer pathways that proceed toward the extracellular side via Asp-96, the Schiff base, Asp-85, and Glu-204 (1–3). Asp-85 is located near the Schiff base and serves as the counterion. The presence of water molecules in this region had been suggested by Fourier transform infrared (FTIR) studies (4, 5), and these were later confirmed as W401, W402, and W406 in the high resolution crystal structure of the BR ground state (PDB code 1C3W; see Fig. 1) (6). The three water molecules form an H-bond network with Asp-85, Arg-82, Asp-212, and the Schiff base. Among the three water molecules, the chemical properties of W402 are of particular interest because the water molecule is located at H-bond distances of 2.63 and 2.87 Å from the carboxyl O atom of Asp-85 and the Schiff base N atom, respectively (see Table 1). From FTIR analysis of various mutants in D2O, the lowest O-D stretching frequency of 2171 cm−1 in BR was assigned to the OW402–H…OAsp-85 bond, implying that this is the strongest H-bond in the network near the Schiff base (7). The band assignment presented in FTIR studies was also supported by previous computational studies by Hayashi et al. (8, 9).
FIGURE 1.
Overview of the BR (PDB code 1C3W) and ASR (PDB code 1XIO) structures. Red spheres indicate O atoms of water molecules. Yellow arrows indicate proton transfer pathways.
TABLE 1.
Comparison of the H-bond geometries near W402 in the BR crystal structures (PDB ID codes 1C3W at 1.55 Å resolution and 2NTU at 1.53 Å resolution) in the QM/MM geometry by Hayashi and Ohmine (8) and in this study (models 1 and 2, Fig. 4)
Distances are in Å, and angles are in degrees. ND, not determined. An underlined number indicates a significant deviation from the distance in the crystal structure. See supplemental S1 for the atomic coordinates of the QM/MM geometries.
Donor…acceptor | 1C3W | 1C3W; QM/MM Hayashi and Ohminea (model 1) | 1C3W; QM/MM (model 1) | 1C3W; QM/MM (model 2) | 2NTU | 2NTU; QM/MM (model 1) | 2NTU; QM/MM (model 2) |
---|---|---|---|---|---|---|---|
Å | Å | Å | Å | Å | Å | Å | |
W402…Asp-85 | 2.63 | 2.58 | 2.61 | 2.57 | 2.52 | 2.63 | 2.59 |
W402…Asp-212 | 2.85 | 4.07 | 2.92 | 2.99 | 3.01 | 2.90 | 2.95 |
W401…Asp-85 | 2.59 | 2.63 | 2.51 | 2.64 | 2.69 | 2.59 | 2.68 |
(OH without acceptor)b | (W401) | (W401) | (W406) | (W401) | (W406) | ||
W406…W401 | 2.75 | 2.77 | 2.76 | 3.06 | 2.69 | 2.78 | 3.07 |
W406…Asp-212 | 2.75 | 2.78 | 2.64 | 2.64 | 2.81 | 2.73 | 2.65 |
Lys-216…W402 | 2.87 | 2.63 | 2.78 | 2.76 | 2.94 | 2.80 | 2.77 |
Arg82…W406 | 2.49 | 2.72 | 2.74 | 2.87 | 2.87 | 2.73 | 2.86 |
Root mean square deviation | ND | ND | 0.19 | 0.18 | ND | 0.17 | 0.18 |
Angle (NLys-216…OW402…OAsp-85) (degrees) | 106 | ND | 108 | 108 | 111 | 111 | 111 |
Water molecule W402 was also found in the crystal structure of Anabaena sensory rhodopsin (ASR) (PDB code 1XIO; see Fig. 1) (10). In contrast to the proton pumping of BR, ASR is believed to function as a photosensor. Despite the high similarity of their protein sequences, some key residues that are functionally important in the proton-pumping event in BR are not conserved in ASR; Asp-96 in BR, which serves as a proton donor to the Schiff base, is replaced with Ser-86 in ASR, and Glu-194 and Glu-204 in BR, which is located in the terminal region of the proton transfer pathway, are replaced with Ser-188 and Asp-198, respectively. Although W402 is present in both BR (6) and ASR (10), Asp-212 in BR is replaced with Pro-206 in ASR (Fig. 1). The absence of W401 and W406 in the corresponding H-bond network of ASR may be associated with the absence of an acidic residue corresponding to Asp-212 in BR. FTIR studies have suggested that the O-D stretch in water molecules was only observed at >2500 cm−1 in ASR (11), which is ∼300 cm−1 higher than the lowest O-D stretching frequency of 2171 cm−1 in BR (7).
The significant difference in the lowest O-D stretching frequency implies that the H-bond properties of W402 are significantly different in BR and ASR. It has been established by FTIR that the lowest O-D stretching frequency of water can be found at less than 2400 cm−1 in a number of proton-pumping rhodopsins (12). The same tendency also holds true for BR and ASR.
In the BR and ASR crystal structures, the angle defined by the Schiff base N, W402 O, and Asp-85 O atoms (NLys…OW402…OAsp) differs significantly; the angle is 106° in BR (6) and 83° in ASR (10). Thus, it was proposed that the angle difference may differentiate the H-bond strength of the water molecules in BR and ASR (11). On the other hand, it also appears that the OW402–H…OAsp angle is more crucial than the NLys…OW402…OAsp angle to the energetics of the H-bond as it involves a H atom. In general, the Odonor–H…Oacceptor angle strongly depends on the Odonor–Oacceptor distance, i.e. the Odonor–H…Oacceptor angle is more linear when the Odonor–Oacceptor bond is shorter (13).
In addition, the NLys…OW402…OAsp angle in the crystal structure can be directly altered in response to changes in the OW402–H…OAsp distance. Because of the nature of H-bonds (14–17), the OW402–H…OAsp-85 distance could also be predominantly determined by the pKa difference between the H-bond donor (W402) and acceptor (Asp) moieties. In general, a smaller pKa difference yields a short, symmetrical H-bond (14–17), and the H atom of OW402–H migrates toward the acceptor moiety, Asp. Unfortunately, the energetics of the OW402–H…OAsp bond, particularly the pKa difference between W402 and the Asp residue, remain unclear. Thus, it is essential to clarify how the OW402–H…OAsp distance is energetically determined in the BR and ASR crystal structures.
Here we present calculated O-D stretching frequencies of D2O near W402 in BR and ASR in the ground state using a large scale quantum mechanical/molecular mechanical (QM/MM) approach. We also present the potential energy profiles of the H-bond between W402 and Asp-85 in BR and between W402 and Asp-75 in ASR and demonstrate that the frequencies and H-bond properties are ultimately determined predominantly by the pKa values of H-bond acceptors Asp-85 and Asp-75. By calculating pKa(Asp-85) in BR and pKa(Asp-75) in ASR through solving the linear Poisson-Boltzmann equation with explicit consideration of the protonation states for all titratable residues, we are finally able to pinpoint the factors that cause a difference of ∼300 cm−1 between the O-D stretching frequencies of W402 in BR and ASR.
EXPERIMENTAL PROCEDURES
As demonstrated in our previous work with photoactive yellow protein (18), we employed the following systematic modeling procedure.
First, we constructed initial molecular models of BR and ASR using their crystal structures and adding hydrogen atoms. Second, to gain better understanding of the electronic structure of W402 and the associated H-bond network, we performed large scale QM/MM calculations for the entire BR and ASR proteins. To gain insight into the QM/MM potential energy profiles of the H-bonds, OW402–H…OAsp-85 in BR and OW402–H…OAsp-75 in ASR, we calculated pKa(Asp-85) in BR and pKa(Asp-75) in ASR by simultaneously titrating all titratable residues in BR and ASR. Technical details of each modeling procedure are summarized below.
Atomic Coordinates and Charges
As a basis for the computations, the crystal structures of BR (PDB codes 1C3W (6) and 2NTU (19)) and ASR (PDB code 1XIO (10)) were used. To generate the initial geometries, the positions of the H atoms were energetically optimized with CHARMM (20) using the CHARMM22 force field. During this procedure, the positions of all non-H atoms were fixed, and the standard charge states of all the titratable groups were maintained (i.e. basic and acidic groups were considered to be protonated and deprotonated, respectively). The Schiff base was considered to be protonated. Atomic partial charges of the amino acids and the Schiff base were adopted from the all-atom CHARMM22 (20) parameter set.
Protonation Pattern and pKa
The present computation is based on the electrostatic continuum model created by solving the linear Poisson-Boltzmann equation with the MEAD program (21). To facilitate a direct comparison with previous computational results (e.g. see Refs. 22 and 23), identical computational conditions and parameters, such as atomic partial charges and dielectric constants, were used. To obtain absolute pKa values of target sites (e.g. pKa(Asp-85) of BR), we calculated the difference in electrostatic energy between the two protonation states, protonated and deprotonated, in a reference model system using a known experimentally measured pKa value (e.g. 4.0 for Asp (24)). The difference in the pKa value of the protein relative to the reference system was added to the known reference pKa value. The experimentally measured pKa values employed as references were 7.2 for the Schiff base (25, 26), 12.0 for Arg, 4.0 for Asp, 9.5 for Cys, 4.4 for Glu, 10.4 for Lys, 9.6 for Tyr (24), and 7.0 and 6.6 for the Nϵ and Nδ atoms of His, respectively (27–29). All other titratable sites were fully equilibrated to the protonation state of the target site during the titration. The ensemble of the protonation patterns was sampled by a Monte Carlo method with Karlsberg (30). The dielectric constants were set to ϵp = 4 inside the protein and ϵw = 80 for water. All computations were performed at 300 K, pH 7.0, and an ionic strength of 100 mm. The linear Poisson-Boltzmann equation was solved using a three-step grid-focusing procedure at resolutions of 2.5, 1.0, and 0.3 Å. The Monte Carlo sampling yielded the probabilities (protonated and deprotonated) of the two protonation states of the molecule. The pKa value was evaluated using the Henderson-Hasselbalch equation. A bias potential was applied to obtain an equal amount of both protonation states (protonated = deprotonated), yielding the pKa value as the resulting bias potential.
QM/MM Calculations
We employed an electrostatic embedding QM/MM scheme and used the Qsite (Version 5.6, Schrödinger, LLC, New York) program code as performed in previous studies (32). We employed the restricted density functional theory method with B3LYP functional and LACVP**+ basis sets. The geometries were refined using a constrained QM/MM optimization whereby the coordinates of the heavy atoms in the surrounding MM region were fixed to the original x-ray coordinates, whereas those of the H atoms in the MM region were optimized with the OPLS-2005 force field. To investigate the energetics of the entire H-bond network, the QM region was defined as follows (Fig. 2): (a) W402 and the residues/groups in the same H-bond network, i.e. W401, W406, Tyr-57, Arg-82, Asp-85, Ser-85, Thr-89, Tyr-185, Asp-212, Lys-216 (the Schiff base), and retinal for BR and (b) W402 and the residues/groups in the same H-bond network, i.e. Tyr-11, Tyr-51, Asp-75, Trp-76, Thr-79, Ser-47, the backbone of Phe-202, Lys-210 (the Schiff base), and retinal for ASR. Other residues that are not in the H-bond network of W402 (e.g. Trp-86 in BR is not involved in the same H-bond network even though the residue corresponds to Trp-76 in ASR) were approximated by the MM force field. The resulting QM/MM optimized geometries are listed in supplemental Table S1 for BR and Table S2 for ASR. The potential energy profile of the H-bond was obtained in the following manner. First, we optimized the geometry without constraints using QM/MM and used the resulting geometry as the initial geometry for the subsequent steps. Next, we moved the H atom from the H-bond donor atom (Odonor) to the acceptor atom (Oacceptor) by 0.05 Å, optimized the geometry by constraining the Odonor–H and Oacceptor…H distances, and then calculated the energy of the resulting geometry. This procedure was repeated until the H atom reached the Oacceptor atom. After obtaining the stable geometry of the QM fragment, we calculated the O-D stretching frequencies of W401, W402, and W406 in BR and of W402 in ASR. The frequency calculations were performed using a numerical differentiation method at the same level as the QM/MM geometry optimization. The calculated frequencies were scaled using a standard factor of 0.9614 for B3LYP (33).
FIGURE 2.
Residues and groups included in the QM regions for BR (a) and ASR (b). Dotted lines indicate hydrogen bonds. c, geometry of moieties near Lys-210 in ASR before (magenta) and after (yellow) the QM/MM optimization is shown.
Calculation of the O-D Stretching Frequencies of D2O from an Empirical Equation
The O-D stretching resulting of D2O were also evaluated on the basis of the resultant QM/MM optimized geometries by using the correlation of stretching frequency with respect to bond distance of the Odonor–D…Oacceptor bond, as previously proposed by Mikenda (34). This correlation was empirically expressed with the O-D stretching frequency, νO–D, and the Odonor–Oacceptor bond distance, rO–O, by
![]() |
where A = 2.11 × 106, and B = 3.23 for the Owater–Dwater…Oacceptor bond (34). Note that these parameters were derived from spectroscopic and x-ray diffraction data for 61 different solid deuterates, e.g. CaSO4·2D2O and MnCl2·6D2O (34).
To predict the frequencies, Equation 1 requires only the Odonor–Oacceptor distance, which can be preferentially taken from the QM/MM optimized geometry. Remarkably, the O-D stretching frequencies calculated by a QM/MM numerical differentiation method were reasonably well reproduced using Equation 1 solely from the optimized Odonor–Oacceptor distance, particularly for frequencies <2500 cm−1 (see Table 3).
TABLE 3.
Model 2, O–D stretching frequencies (in cm−1) of the water molecules near W402 in BR
The QM/MM results of model 2 on the basis of the 1C3W structure primarily discussed in this study.
Donor…acceptor | FTIR-2a | 1C3W |
2NTU |
||
---|---|---|---|---|---|
QM/MM (model 2) | Empirical equationb (model 2) | QM/MM (model 2) | Empirical equationb (model 2) | ||
W402…Asp-85 | 2171(lowest) | 2078 (lowest) | 2201 (lowest) | 2120 (lowest) | 2240 (lowest) |
W402…Asp-212 | 2599 | 2690 | 2592 | 2670 | 2572 |
W401…Asp-85 | 2323 | 2400 | 2311 | 2411 | 2359 |
W406…(ND) | 2690 | 2678 | 2727 | 2689 | 2727 |
W406…W401 | 2636 | 2658 | 2620 | 2656 | 2622 |
W406…Asp-212 | 2292 | 2312 | 2303 | 2302 | 2328 |
R2 to FTIRc | 0.94c | 0.99c | 0.96c | 0.98c |
Notably, Equation 1 indicates that the O-D stretching frequency of 2500 cm−1 corresponds to the Oacceptor–Odonor distance of ∼2.8 Å. The deviation from the values calculated with the QM/MM numerical differentiation method appears to be more pronounced for frequencies >2500 cm−1 (Fig. 3). Because the original parameters were derived from solid hydrates (34), the larger deviation for weak H-bonds at these frequencies may be due to H-bond patterns or mode couplings specific to the protein environments.
FIGURE 3.
The geometric correlation of the O-D stretching frequencies with the Oacceptor…H-Odonor bond of H2O (correlation with the Odonor–Oacceptor distance). Each open square (□) represents the Odonor–Oacceptor distance of the QM/MM optimized geometries (Tables 1 and 4) and the O-D stretching frequency calculated by a QM/MM numerical differentiation method (Tables 2, 3, and 5). The solid curve was determined using an empirical equation (Equation 1) proposed by Mikenda (34).
RESULTS
H-bond Geometries Near W402 in BR
The O-D stretching frequencies of D2O predominantly depend on the H-bond geometry, particularly the H-bond donor-acceptor distance (Odonor–Oacceptor) (34). If the assumed H-bond pattern is not correct, the resulting QM/MM optimized geometries significantly deviate from the original crystal structure. Thus, although the QM/MM optimized geometries (especially the heavy atom positions) are not always consistent with the crystal structures, it is necessary to carefully evaluate the H-bond pattern with respect to the atomic coordinates of the crystal structures before frequency calculations.
In BR, two H-bond patterns are geometrically possible for the H-bond network of W402 (Fig. 4). A model previously presented from FTIR studies (7) suggests that an H atom of W401 has no H-bond acceptor (model 1), whereas an alternative model suggests that an H atom of W406 has no H-bond acceptor (model 2). The QM/MM-optimized geometries of the two models appear to represent the actual H-bond patterns of the crystal structures because of the low root mean square deviation (0.17–0.19 Å, Table 1), implying that the two models probably coexist as tautomers in the actual BR protein environment.
FIGURE 4.
H-bond geometries of water molecules near the Schiff base in BR and ASR. Pink arrows in model 1 indicate mode couplings of the two O-D bonds. Red dotted lines indicate “strong H-bonds” reported in FTIR studies (7).
Model 1 resembles the QM/MM geometry reported by Hayashi and Ohmine (8). However, in their QM/MM geometry, the OW402–OAsp-212 distance was significantly elongated to 4.07 Å, and the H-bond is absent (8). This varies significantly from the crystal structures and the present QM/MM geometries (OW402–OAsp-212 = 2.85–3.01 Å, Tables 2 and 3). In the reported QM/MM geometry, W402 forms a new H-bond with another carboxyl O atom of Asp-212 (Fig. 4). Hence, it is to be considered that the band assignment reported by Hayashi and Ohmine (8) refers to another possible H-bond geometry that differs from the crystal structures.
TABLE 2.
Model 1, O-D stretching frequencies (in cm−1) of the water molecules near W402 in BR
Connecting lines indicate the presence of coupling between the two modes. n.d., not determined; R2, correlation coefficient.
The QM/MM geometries imply that model 1 can explain the Odonor–Oacceptor distances in the crystal structure provided by Luecke et al. (PDB code 1C3W) (6) more reasonably than model 2 (Table 1). On the other hand, model 2 appears to explain the Odonor–Oacceptor distances in the crystal structure determined by Lanyi and Schobert (PDB code 2NTU) (19) more reasonably, namely in terms of OW402–OAsp-85, OW402–OAsp-212, OW401–OAsp-85, and OArg-82–OW406 distances (Table 1). However, the small discrepancies could also be due to being artifacts of the crystallization process. Because the resolutions of the two crystal structures are 1.53 and 1.55 Å (which are essentially the same), both of the H-bond tautomers are likely to contribute to the O-D stretching frequencies measured by FTIR. If both tautomers are present, this might corroborate the rapid H/D exchange observed in this region.
Calculated O-D Stretching Frequencies in BR
In FTIR studies (7), the three H-bonds OW401–H…OAsp-85, OW402–H…OAsp-85, and OW406–H…OAsp-212 were classified as being strong H-bonds because of the assignment of low frequencies (Fig. 4).
Model 1
In FTIR studies (7) the lowest frequency of 2171 cm−1 was assigned to OW402–H…OAsp-85. However, in model 1, OW402–H…OAsp-85 was longer than OW401–H…OAsp-85 (Table 1), leading to a larger calculated O-D stretching frequency for OW402–H…OAsp-85 than OW401–H…OAsp-85 (Table 2), as predicted from the empirical equation by Mikenda (Equation 1) (34).
In model 1, two mode couplings between OW401–H…OAsp-85 and OW401–H…OAsp-85 and between OW406–H…OAsp-212 and OW406–H…OW401 were observed. However, the latter coupling is not consistent with the FTIR assignments. In FTIR studies, H-bonds OW406–H…OAsp-212 and OW406–H…OW401 were assigned as strong (2292 cm−1) and weak (2599 cm−1), respectively (Table 2) (7). Because mode couplings are unlikely to occur between a strong and weak H-bond, model 1 fails to sufficiently explain the FTIR assignments.
Model 2
In contrast to model 1, model 2 can reasonably explain the strong H-bond (2292 cm−1) of OW406–H…OAsp-212 and the weak H-bond (2636 cm−1) of OW406–H…OW401 that were assigned by FTIR studies (Table 3). Furthermore, the lowest calculated frequency was observed for OW402–H…OAsp-85, which is also consistent with FTIR studies (7). Significantly high correlations of the frequencies predicted from Equation 1 suggest that assignment of the O-D stretching frequencies on the basis of model 2 is also justified by the H-bond geometry. Note that no remarkable mode couplings were observed for model 2 in the QM/MM calculations.
In the following section further analysis of BR will be discussed on the basis of model 2 using the crystal structure by Luecke et al. (PDB code 1C3W) (6) unless otherwise specified.
H-bond Geometries near W402 and O-D Stretching Frequencies in ASR
In contrast to BR, the ASR crystal structure only possesses a single water molecule, W402, near the Schiff base (Fig. 4). The H-bond geometry of the crystal structure was reasonably reproduced by the QM/MM geometry optimization (Table 4). In the QM/MM calculations, the lowest O-D stretching frequency of 2376 cm−1 in ASR, which was found for OW402–H…OAsp-75 (Table 5), was significantly larger (by ∼300 cm−1) than that for the corresponding OW402–H…OAsp-85 in BR (2078 cm−1, Table 3). In FTIR studies, the O-D stretching of D2O was only observed in the >2500-cm−1 region for ASR (11), which was also ∼300 cm−1 higher than the lowest frequency of 2171 cm−1 in BR (7). Because the main purpose of this study is to understand the reason for the significant discrepancy between the lowest frequencies of D2O (assumed as W402 (11)) of BR and ASR, these computational results are sufficiently accurate for our purpose to describe the chemical properties of W402.
TABLE 4.
H-bond geometries near W402 in the ASR crystal structure (PDB ID code 1XIO at 2.00 Å resolution) and the QM/MM geometry
ND, not determined. Seesupplemental S1 for the atomic coordinates of the QM/MM geometries. Distances are in Å, and angles are in degrees.
Donor…acceptor | 1XIO | QM/MM, deprotonated Asp-75 |
---|---|---|
Å | Å | |
W402…Asp-75 | 2.74 | 2.71 |
W402…Tyr-51 | 2.92 | 2.97 |
Trp-76…W402 | 2.94 | 2.84 |
Lys-210…W402 | 3.02 | 3.00 |
Root mean square deviationa | ND | 0.17 |
Angle (NLys-210…OW402…OAsp-75) (degrees) | 83 | 92 |
a Excluding Lys-210 due to obvious displacement of the sidechain carbon atoms from the original crystal structure (Fig. 2c).
TABLE 5.
O–D stretching frequencies (in cm−1) of W402 in ASR
Donor…acceptor | FTIRa | Deprotonated Asp-75 |
|
---|---|---|---|
QM/MM | Empirical equation | ||
W402…Asp-75 | >2500 | 2376 | 2392 |
W402…Tyr-51 | >2500 | 2573 | 2583 |
a See Ref. 11.
DISCUSSION
Significant Differences between the H-bond Energy Profiles of BR and ASR
To identify the origin of the difference in O-D stretching frequencies of W402, we analyzed the potential energy profiles of H-bonds OW402–H…OAsp-85 in BR and OW402–H…OAsp-75 in ASR by altering the H atom position along the OW402–OAsp-85/75 bond (see Fig. 5). In both H-bonds, the energy minimum was found at the W402 moiety rather than at the Asp moiety, suggesting that in the ground state W402 exists as H2O in the presence of deprotonated Asp in BR and ASR.
FIGURE 5.
Potential energy profiles along the proton transfer coordinates of OW402–H…OAsp-85 in BR (black solid line) and OW402–H…OAsp-75 in ASR (blue solid line). ΔE describes the difference in energy relative to the energy minimum. Boxed arrows indicate the shifts in the potential energy curve accompanied by the pKa(Asp) change from ASR to BR. The marginal energy drop near 1.74 Å in OW402–H of ASR was due to an alteration of the H-bond pattern; below 1.74 Å, Ser-47 donates an H-bond to the carboxyl O atom of Asp-75, which is simultaneously H-bonded by W402. At 1.74 Å, the hydroxyl H atom is oriented toward another carboxyl O atom of Asp-75 due to the unusual proximity of the H atom of W402 to Asp-75.
On the other hand, the potential energy profiles of the Asp moiety were significantly different for BR and ASR. For an H atom along the OW402–OAsp-85/75 bond, the energy of the Asp-85 moiety of BR is significantly lower than that of the Asp-75 moiety of ASR. Thus, it is more favorable for an H atom of OW402–H to be at the Asp moiety in BR than in ASR, indicating that pKa(Asp) is significantly higher for BR than ASR (arrow 1 in Fig. 5). As clearly shown in Fig. 5, the lower energy near the Asp-85 moiety in BR leads to greater broadening of the potential-well width near W402 toward Asp-85, which corresponds to an elongation of OW402–H. Thus, the H atom of OW402–H migrates toward the H-bond acceptor Asp more in BR than in ASR (arrow 2 in Fig. 5). It should be noted that the longer OW402–H corresponds to a smaller O-D stretching frequency. It is also obvious from the width of the entire well of the potential energy profiles that the OW402–OAsp-85 length in BR (∼2.6 Å, Table 1), which is shorter than the OW402–OAsp-75 length in ASR (∼2.7 Å, Table 4), is a result of the smaller pKa difference between W402 and the Asp residue in BR relative to the corresponding difference in ASR (arrow 3 in Fig. 5). Thus, from the potential energy profiles (Fig. 5), the difference of ∼300 cm−1 in the O-D stretching frequency is mainly due to the difference between the pKa values for the Asp residues in BR and ASR.
Factors that Differentiate pKa(Asp) of BR and ASR
By solving the linear Poisson-Boltzmann equation, we calculated pKa(Asp-85) in BR and pKa(Asp-75) in ASR to be 1.5 and −5.1, respectively, in the presence of the protonated Schiff base (Table 6). The calculated pKa(Asp-85) of 1.5 is consistent with an experimentally measured pKa(Asp-85) of 2.2–2.6 in the presence of the protonated Schiff base (2, 35). Although pKa(Asp-75) is not known, the significantly low pKa(Asp-85) in BR relative to pKa(Asp-75) in ASR is consistent with the QM/MM-calculated potential energy profiles (Fig. 5). The significantly low pKa(Asp-75) in ASR appears to be consistent with the fact that Asp-75 remains deprotonated throughout the photocycle (36).
TABLE 6.
Contribution of residues to the pKa shifts of Asp-85 in BR and Asp-75 in ASR (relative to the bulk solvent) in the presence of the protonated Schiff base (in pKa units)
For clarity, residue pairs that are responsible for differences of <0.7 between pKa(Asp-85) and pKa(Asp-75) are not listed except for the Glu-204/Asp-198 pair. Side, side chain; bb, backbone.
BR (1C3W) | pKa(Asp-85) = 1.5 |
ASR (1XIO) | pKa (Asp-75) = −5.1 |
Difference total | ||||
---|---|---|---|---|---|---|---|---|
Side | bb | Total | Side | bb | Total | |||
Ile-52 | −0.1 | 0.0 | −0.1 | Trp-46 | −0.6 | −0.3 | −0.9 | 0.8 |
Ala-53 | 0.0 | −0.5 | −0.5 | Ser-47 | −2.6 | −0.6 | −3.2 | 2.7 |
Tyr-57 | −0.5 | −0.4 | −0.9 | Tyr-51 | 0.2 | −0.4 | −0.2 | −0.7 |
Arg-82 | −3.9 | 0.8 | −3.1 | Arg72 | −2.1 | 0.9 | −1.2 | −1.9 |
Glu-194 | 1.2 | −0.1 | 1.1 | Ser-188 | 0.0 | 0.0 | 0.0 | 1.1 |
Glu-204 | 0.0 | 0.1 | 0.1 | Asp-198 | 0.6 | 0.0 | 0.6 | −0.5 |
Asp-212 | 6.3 | −0.1 | 6.2 | Pro-206 | −0.1 | 6.3 | ||
Schiff | −9.3 | Schiff | −10.3 | 1.0 |
Asp-212 in BR
The most significant difference observed between pKa(Asp-85) and pKa(Asp-75) was induced by the substitution of Asp-212 in BR for Pro-206 in ASR (Fig. 6), which is responsible for the pKa difference of ∼6 (Table 6). Because Asp-212 is an H-bond acceptor for OW402–H and is thus far the most proximal negative charge to Asp-85 (5.1 Å), Asp-212 upshifts pKa(Asp-85) by the largest amount among all residues in BR (Table 6).
FIGURE 6.
Residues that contribute to differences between pKa(Asp) of BR (a) and ASR (b).
It is widely known that both Asp-85 and Asp-212 are ionized in the ground state (2, 37, 38). However, the deprotonated state of Asp-212 is more energetically stable than the deprotonated state of Asp-85 as seen in the lower (pKa(Asp-212) = −2.0 and pKa(Asp-85) = 1.5) (Table 7); this is consistent with previous pKa assignments of <1 for Asp-212 (39, 40) or pKa(Asp-85) < pKa(Asp-212) (41, 42). The deprotonated state of Asp-212 was facilitated by H-bond donations from Tyr-57 (decreasing pKa(Asp-212) by 3.4) and Tyr-185 (decreasing pKa(Asp-212) by 3.3). In addition, Arg-82 is closer to Asp-212 (3.8 Å) than Asp-85 (6.6 Å) (Fig. 6), stabilizing the ionized state of Asp-212 more effectively than Asp-85 (increasing pKa by ∼6, Table 7).
TABLE 7.
Key residues that alter the pKa difference between Asp-85 and Asp-212 in the presence of the protonated Schiff base (in pKa units)
Side, side chain; bb, backbone; ND = not determined.
BR (1C3W) residue | pKa (Asp-85) = 1.5 |
pKa (Asp-212) = −2.0 |
||||
---|---|---|---|---|---|---|
Side | bb | Total | Side | bb | Total | |
Met-56 | −0.4 | −0.8 | −1.2 | −0.1 | −0.3 | −0.3 |
Tyr-57 | −0.5 | −0.3 | −0.9 | v3.2 | −0.2 | −3.4 |
Ala-81 | 0.0 | 0.7 | 0.8 | 0.0 | 0.1 | 0.1 |
Arg-82 | −3.9 | 0.8 | −3.1 | −6.4 | 0.1 | −6.3 |
Asp-85 | ND | ND | ND | 6.4 | −0.4 | 6.0 |
Trp-86 | −0.2 | −0.6 | −0.8 | v0.8 | −0.3 | −1.0 |
Thr-89 | −2.0 | −0.9 | −2.9 | 0.1 | −0.2 | −0.1 |
Tyr-185 | −0.3 | −0.1 | −0.4 | −3.2 | −0.1 | −3.3 |
Glu-194 | 1.2 | −0.1 | 1.2 | 1.8 | −0.1 | 1.7 |
Phe-208 | 0.0 | 0.2 | 0.2 | 0.0 | 1.8 | 1.8 |
Asp-212 | 6.3 | −0.1 | 6.2 | ND | ND | ND |
Schiff | −9.5 | 0.2 | −9.3 | −9.3 | 0.6 | −8.7 |
Although the most significant difference between pKa(Asp-85) and pKa(Asp-75) induced by the substitution of Asp-212 in BR for Pro-206 in ASR is remarkable, this is not sufficient to entirely explain the difference between pKa(Asp-85) and pKa(Asp-75) (see below), which is in agreement with mutational studies of ASR mutant P206D (43).
Ser-47 in ASR
Ser-47 donates an H-bond to Asp-75 in ASR, decreasing pKa(Asp-75) by 3.2, whereas the corresponding residue in BR is the non-polar Ala-53 (Table 6). This was responsible for a pKa difference of 2.7 between Asp-85 and Asp-75 (Table 6).
Glu-194–Glu-204 Pair in BR
Another remarkable pKa(Asp-85) upshift in BR originates from the Glu-194–Glu-204 moiety near the terminal region of the proton transfer pathway (2, 3) (Fig. 6). It is likely that the two acidic residues share a proton. The corresponding acidic residue pair is absent in ASR, thereby contributing slightly to the larger pKa(Asp) in BR relative to that in ASR (∼0.5 pKa units, Table 6). In this study Glu-204 is protonated and Glu-194 is deprotonated when Ser-193 donates an H-bond to Glu-194, which is consistent with the previously reported pKa(Glu-204) of ∼9 (35, 44, 45).
Nevertheless, the geometry of the crystal structure also suggests that protonated Glu-194 and deprotonated Glu-204 are almost equally possible when Ser-193 donates an H-bond to Glu-204, as suggested by Gerwert and co-workers (46). Thus, it would be more plausible to consider the two acidic residues as a pair that possesses ∼1 H+ as concluded in previous QM/MM simulations (47). This resembles a pair of acidic residues, Glu-L212 and Asp-L213, in the proton transfer pathway of photosynthetic reaction centers, which has been interpreted as sharing ∼1 H+ (48–50). Deprotonated Glu-194 contributes to an increase in pKa(Asp-85) of ∼1 (Table 6); this, therefore, corroborates the linkage between pKa(Asp-85) and pKa(Glu-204) previously reported in mutant BR studies (35) if we consider Glu-194–Glu-204 as a pair of acidic residues sharing ∼1 H+.
Orientation of Arg-82 in BR and Arg-72 ASR
Arg-82 was found to contribute to an increase in pKa(Asp-85) of 3.1 in BR (Table 6), which is in agreement with a similar increase (∼4.5) reported for BR mutants R82Q and R82A (51, 52). Irrespective of the conservation of Arg-82/Arg-72 in BR/ASR, their influence on pKa(Asp-85) and pKa(Asp-75) significantly differs by ∼2 because the side chain of Arg-72 is oriented away from the Schiff base and toward the extracellular side in ASR (OAsp-75–NArg-72 = 9.6 Å) (10), which is the opposite of the orientation of Arg-82 in BR (OAsp-85–NArg-82 = 6.6 Å) (6) (Fig. 6). Thus, the differing orientations of the Arg side chains decrease the pKa(Asp) difference between BR and ASR (Table 7).
In summary, the significantly large pKa(Asp-85) in BR relative to pKa(Asp-75) in ASR is largely due to the presence of Asp-212 near Asp-85 in BR and the absence of the corresponding acidic residue in ASR. In addition, the donation of an H-bond from Ser-47 to Asp-75 lowers the pKa(Asp-75) in ASR by facilitating deprotonation. The corresponding polar residue is absent in BR; this upshifts pKa(Asp-85) in BR relative to pKa(Asp-75) in ASR. The Glu-194–Glu-204 pair is located near the terminal region of the proton transfer pathway in BR, whereas only a single acidic residue is located near the corresponding position in ASR. This also upshifts pKa(Asp-85) in BR relative to pKa(Asp-75) in ASR.
Conclusions
QM/MM calculations revealed that the heavy atom positions in the BR crystal structures could be explained by one of two models. Model 2 can reasonably explain the strong H-bond (2292 cm−1) of OW406–H…OAsp-212 and the weak H-bond (2636 cm−1) of OW406–H…OW401, as previously assigned by FTIR (Table 3). In model 2, the lowest calculated frequency was observed at OW402–H…OAsp-85, which is also consistent with FTIR studies. Essentially the same frequencies were also predicted solely from the Odonor–Oacceptor distances using Equation 1 proposed by Mikenda (34). This suggests that Equation 1 can be used to support the frequency assignments made from spectroscopic studies. In ASR, the lowest calculated O-D stretching frequency is observed at OW402–H…OAsp-75, which is ∼300 cm−1 higher than that of BR (Tables 3 and 6), as previously reported in FTIR studies of BR (7) and ASR (11).
The potential energy profiles of two H-bonds, OW402–H…OAsp-85 in BR and OW402–H…OAsp-75 in ASR, demonstrate that the significant pKa difference between Asp-85 in BR and Asp-75 in ASR is ultimately responsible for the difference of ∼300 cm−1 in the O-D stretching frequency of W402 between BR and ASR (Fig. 5). Electrostatic calculations resulted in a pKa(Asp-85) of 1.5 in BR and a pKa(Asp-75) of −5.1 in ASR (Table 6). The pKa difference is mainly due to differences in the influences of Ala-53, Arg-82, Glu-194, and Asp-212 on Asp-85 in BR and the corresponding residues Ser-47, Arg-72, Ser-188, and Pro-206 on Asp-75 in ASR, indicating the presence of long range electrostatic interactions along the proton transfer pathways (31). A strongly H-bonded water near the Schiff base, which results from electrostatic interactions between Asp-85 and these acidic residues in BR, could play a key role in proton transfer.
This work was supported by the JST PRESTO program (to H. I.), Grants-in-aid for Scientific Research from the Ministry of Education, Culture, Sports, Science, and Technology (MEXT) of Japan 22740276 (to K. S.) and 20108014 (to H. K.), Special Coordination Fund for Promoting Science and Technology of MEXT (to H. I.), Takeda Science Foundation (to H. I.), Kyoto University Step-up Grant-in-aid for Young Scientists (to H. I.), and Grant for Basic Science Research Projects from the Sumitomo Foundation (to H. I.).

This article contains supplemental Tables S1 and S2.
- BR
- bacteriorhodopsin
- ASR
- Anabaena sensory rhodopsin
- QM/MM
- quantum mechanical/molecular mechanical.
REFERENCES
- 1. Lanyi J. K. (1998) Understanding structure and function in the light-driven proton pump bacteriorhodopsin. J. Struct. Biol. 124, 164–178 [DOI] [PubMed] [Google Scholar]
- 2. Balashov S. P. (2000) Protonation reactions and their coupling in bacteriorhodopsin. Biochim. Biophys. Acta 1460, 75–94 [DOI] [PubMed] [Google Scholar]
- 3. Kandori H. (2000) Role of internal water molecules in bacteriorhodopsin. Biochim. Biophys. Acta 1460, 177–191 [DOI] [PubMed] [Google Scholar]
- 4. Maeda A., Sasaki J., Yamazaki Y., Needleman R., Lanyi J. K. (1994) Interaction of aspartate-85 with a water molecule and the protonated Schiff base in the L intermediate of bacteriorhodopsin. A Fourier-transform infrared spectroscopic study. Biochemistry 33, 1713–1717 [DOI] [PubMed] [Google Scholar]
- 5. Kandori H., Yamazaki Y., Sasaki J., Needleman R., Lanyi J. K., Maeda A. (1995) Water-mediated proton transfer in proteins. An FTIR study of bacteriorhodopsin. J. Am. Chem. Soc. 117, 2118–2119 [Google Scholar]
- 6. Luecke H., Schobert B., Richter H. T., Cartailler J. P., Lanyi J. K. (1999) Structure of bacteriorhodopsin at 1.55 A resolution. J. Mol. Biol. 291, 899–911 [DOI] [PubMed] [Google Scholar]
- 7. Shibata M., Kandori H. (2005) FTIR studies of internal water molecules in the Schiff base region of bacteriorhodopsin. Biochemistry 44, 7406–7413 [DOI] [PubMed] [Google Scholar]
- 8. Hayashi S., Ohmine I. (2000) Proton transfer in bacteriorhodopsin. Structure, excitation, IR spectra, and potential energy surface analyses by an ab initio QM/MM method. J. Phys. Chem. B. 104, 10678–10691 [Google Scholar]
- 9. Hayashi S., Tajkhorshid E., Kandori H., Schulten K. (2004) Role of hydrogen-bond network in energy storage of bacteriorhodopsin light-driven proton pump revealed by ab initio normal-mode analysis. J. Am. Chem. Soc. 126, 10516–10517 [DOI] [PubMed] [Google Scholar]
- 10. Vogeley L., Sineshchekov O. A., Trivedi V. D., Sasaki J., Spudich J. L., Luecke H. (2004) Anabaena sensory rhodopsin. A photochromic color sensor at 2.0 Å. Science 306, 1390–1393 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11. Furutani Y., Kawanabe A., Jung K. H., Kandori H. (2005) FTIR spectroscopy of the all-trans form of Anabaena sensory rhodopsin at 77 K. Hydrogen bond of a water between the Schiff base and Asp-75. Biochemistry 44, 12287–12296 [DOI] [PubMed] [Google Scholar]
- 12. Kandori H. (2010) in Hydrogen Bonding and Transfer in the Excited State (Han K. L., Zhao G. J., eds) pp. 377–391, John-Wiley & Sons Ltd., West Sussex, UK [Google Scholar]
- 13. Frey P. A. (2006) in Isotope Effects in Chemistry and Biology (Kohen A., Limbach H. H., eds) pp. 975–993, CRC Press, Boca Raton, FL [Google Scholar]
- 14. Cleland W. W., Kreevoy M. M. (1994) Low barrier hydrogen bonds and enzymic catalysis. Science 264, 1887–1890 [DOI] [PubMed] [Google Scholar]
- 15. Frey P. A., Whitt S. A., Tobin J. B. (1994) A low barrier hydrogen bond in the catalytic triad of serine proteases. Science 264, 1927–1930 [DOI] [PubMed] [Google Scholar]
- 16. Perrin C. L., Nielson J. B. (1997) “Strong” hydrogen bonds in chemistry and biology. Annu. Rev. Phys. Chem. 48, 511–544 [DOI] [PubMed] [Google Scholar]
- 17. Schutz C. N., Warshel A. (2004) The low barrier hydrogen bond (LBHB) proposal revisited. The case of the Asp. His pair in serine proteases. Proteins 55, 711–723 [DOI] [PubMed] [Google Scholar]
- 18. Saito K., Ishikita H. (2012) Energetics of short hydrogen bonds in photoactive yellow protein. Proc. Natl. Acad. Sci. U.S.A. 109, 167–172 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19. Lanyi J. K., Schobert B. (2007) Structural changes in the L photointermediate of bacteriorhodopsin. J. Mol. Biol. 365, 1379–1392 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20. Brooks B. R., Bruccoleri R. E., Olafson B. D., States D. J., Swaminathan S., Karplus M. (1983) CHARMM. A program for macromolecular energy minimization and dynamics calculations. J. Comput. Chem. 4, 187–217 [Google Scholar]
- 21. Bashford D., Karplus M. (1990) pKa values of ionizable groups in proteins. Atomic detail from a continuum electrostatic model. Biochemistry 29, 10219–10225 [DOI] [PubMed] [Google Scholar]
- 22. Ishikita H., Knapp E. W. (2007) Protonation states of ammonia/ammonium in the hydrophobic pore of ammonia transporter protein AmtB. J. Am. Chem. Soc. 129, 1210–1215 [DOI] [PubMed] [Google Scholar]
- 23. Ishikita H. (2010) Origin of the pKa shift of the catalytic lysine in acetoacetate decarboxylase. FEBS Lett. 584, 3464–3468 [DOI] [PubMed] [Google Scholar]
- 24. Nozaki Y., Tanford C. (1967) Acid-base titrations in concentrated guanidine hydrochloride. Dissociation constants of the guamidinium ion and of some amino acids. J. Am. Chem. Soc. 89, 736–742 [DOI] [PubMed] [Google Scholar]
- 25. Bassov T., Sheves M. (1986) Alteration of pKa of the bacteriorhodopsin protonated Schiff base. A study with model compounds. Biochemistry 25, 5249–5258 [Google Scholar]
- 26. Rousso I., Friedman N., Sheves M., Ottolenghi M. (1995) pKa of the protonated Schiff base and aspartic 85 in the bacteriorhodopsin binding site is controlled by a specific geometry between the two residues. Biochemistry 34, 12059–12065 [DOI] [PubMed] [Google Scholar]
- 27. Tanokura M. (1983) 1H-NMR study on the tautomerism of the imidazole ring of histidine residues. I. Microscopic pK values and molar ratios of tautomers in histidine-containing peptides. Biochim. Biophys. Acta 742, 576–585 [DOI] [PubMed] [Google Scholar]
- 28. Tanokura M. (1983) 1H NMR study on the tautomerism of the imidazole ring of histidine residues. II. Microenvironments of histidine 12 and histidine 119 of bovine pancreatic ribonuclease A. Biochim. Biophys. Acta 742, 586–596 [DOI] [PubMed] [Google Scholar]
- 29. Tanokura M. (1983) 1H nuclear magnetic resonance titration curves and microenvironments of aromatic residues in bovine pancreatic ribonuclease A. J. Biochem. 94, 51–62 [DOI] [PubMed] [Google Scholar]
- 30. Rabenstein B., Knapp E. W. (2001) Calculated pH-dependent population and protonation of carbon-monoxymyoglobin conformers. Biophys. J. 80, 1141–1150 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31. Calimet N., Ullmann G. M. (2004) The influence of a transmembrane pH gradient on protonation probabilities of bacteriorhodopsin. The structural basis of the back-pressure effect. J. Mol. Biol. 339, 571–589 [DOI] [PubMed] [Google Scholar]
- 32. Saito K., Ishida T., Sugiura M., Kawakami K., Umena Y., Kamiya N., Shen J. R., Ishikita H. (2011) Distribution of the cationic state over the chlorophyll pair of photosystem II reaction center. J. Am. Chem. Soc. 133, 14379–14388 [DOI] [PubMed] [Google Scholar]
- 33. Scott A. P., Radom L. (1996) Harmonic vibrational frequencies. An evaluation of Hartree-Fock, Møller-Plesset, quadratic configuration interaction, density functional theory, and semiempirical scale factors. J. Phys. Chem. 100, 16502–16513 [Google Scholar]
- 34. Mikenda W. (1986) Stretching frequency versus bond distance correlation of O-D(H) … Y (Y = N, O, S, Se, Cl, Br, I) hydrogen bonds in solid hydrates. J. Mol. Struct. 147, 1–15 [Google Scholar]
- 35. Richter H. T., Brown L. S., Needleman R., Lanyi J. K. (1996) A linkage of the pKa values of Asp-85 and Glu-204 forms part of the reprotonation switch of bacteriorhodopsin. Biochemistry 35, 4054–4062 [DOI] [PubMed] [Google Scholar]
- 36. Bergo V. B., Ntefidou M., Trivedi V. D., Amsden J. J., Kralj J. M., Rothschild K. J., Spudich J. L. (2006) Conformational changes in the photocycle of Anabaena sensory rhodopsin. Absence of the Schiff base counterion protonation signal. J. Biol. Chem. 281, 15208–15214 [DOI] [PubMed] [Google Scholar]
- 37. Song Y., Mao J., Gunner M. R. (2003) Calculation of proton transfers in bacteriorhodopsin bR and M intermediates. Biochemistry 42, 9875–9888 [DOI] [PubMed] [Google Scholar]
- 38. Bombarda E., Becker T., Ullmann G. M. (2006) Influence of the membrane potential on the protonation of bacteriorhodopsin. Insights from electrostatic calculations into the regulation of proton pumping. J. Am. Chem. Soc. 128, 12129–12139 [DOI] [PubMed] [Google Scholar]
- 39. Metz G., Siebert F., Engelhard M. (1992) Asp-85 is the only internal aspartic acid that gets protonated in the M intermediate and the purple-to-blue transition of bacteriorhodopsin. A solid-state 13C CP-MAS NMR investigation. FEBS Lett. 303, 237–241 [DOI] [PubMed] [Google Scholar]
- 40. Balashov S. P., Govindjee R., Imasheva E. S., Misra S., Ebrey T. G., Feng Y., Crouch R. K., Menick D. R. (1995) The two pKa values of aspartate 85 and control of thermal isomerization and proton release in the arginine 82 to lysine mutant of bacteriorhodopsin. Biochemistry 34, 8820–8834 [DOI] [PubMed] [Google Scholar]
- 41. Bashford D., Gerwert K. (1992) Electrostatic calculations of the pKa values of ionizable groups in bacteriorhodopsin. J. Mol. Biol. 224, 473–486 [DOI] [PubMed] [Google Scholar]
- 42. Sampogna R. V., Honig B. (1994) Environmental effects on the protonation states of active site residues in bacteriorhodopsin. Biophys. J. 66, 1341–1352 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43. Choi A. R., Kim S. Y., Yoon S. R., Bae K., Jung K. H. (2007) Substitution of Pro-206 and Ser-86 residues in the retinal binding pocket of Anabaena sensory rhodopsin is not sufficient for proton pumping function. J. Microbiol. Biotechnol. 17, 138–145 [PubMed] [Google Scholar]
- 44. Balashov S. P., Imasheva E. S., Govindjee R., Ebrey T. G. (1996) Titration of aspartate-85 in bacteriorhodopsin. What it says about chromophore isomerization and proton release. Biophys. J. 70, 473–481 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45. Dioumaev A. K., Richter H. T., Brown L. S., Tanio M., Tuzi S., Saito H., Kimura Y., Needleman R., Lanyi J. K. (1998) Existence of a proton transfer chain in bacteriorhodopsin. Participation of Glu-194 in the release of protons to the extracellular surface. Biochemistry 37, 2496–2506 [DOI] [PubMed] [Google Scholar]
- 46. Rammelsberg R., Huhn G., Lübben M., Gerwert K. (1998) Bacteriorhodopsin intramolecular proton-release pathway consists of a hydrogen-bonded network. Biochemistry 37, 5001–5009 [DOI] [PubMed] [Google Scholar]
- 47. Goyal P., Ghosh N., Phatak P., Clemens M., Gaus M., Elstner M., Cui Q. (2011) Proton storage site in bacteriorhodopsin. New insights from quantum mechanics/molecular mechanics simulations of microscopic pKa and infrared spectra. J. Am. Chem. Soc. 133, 14981–14997 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48. Alexov E. G., Gunner M. R. (1999) Calculated protein and proton motions coupled to electron transfer. Electron transfer from QA− to QB in bacterial photosynthetic reaction centers. Biochemistry 38, 8253–8270 [DOI] [PubMed] [Google Scholar]
- 49. Gerencsér L., Maróti P. (2001) Retardation of proton transfer caused by binding of the transition metal ion to the bacterial reaction center is due to pKa shifts of key protonatable residues. Biochemistry 40, 1850–1860 [DOI] [PubMed] [Google Scholar]
- 50. Ishikita H., Morra G., Knapp E. W. (2003) Redox potential of quinones in photosynthetic reaction centers from Rhodobacter sphaeroides. Dependence on protonation of Glu-L212 and Asp-L213. Biochemistry 42, 3882–3892 [DOI] [PubMed] [Google Scholar]
- 51. Otto H., Marti T., Holz M., Mogi T., Stern L. J., Engel F., Khorana H. G., Heyn M. P. (1990) Substitution of amino acids Asp-85, Asp-212, and Arg-82 in bacteriorhodopsin affects the proton release phase of the pump and the pK of the Schiff base. Proc. Natl. Acad. Sci. U.S.A. 87, 1018–1022 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52. Brown L. S., Bonet L., Needleman R., Lanyi J. K. (1993) Estimated acid dissociation constants of the Schiff base, Asp-85, and Arg-82 during the bacteriorhodopsin photocycle. Biophys. J. 65, 124–130 [DOI] [PMC free article] [PubMed] [Google Scholar]