Full Text (PDF) [PDF]

Aug 5, 2014 - propionate D, perpendicular to the membrane surface (Fig. S1), which is a key step of the ... (section 2)

1 downloads 5 Views 1MB Size

Recommend Stories


Full text (PDF)
I cannot do all the good that the world needs, but the world needs all the good that I can do. Jana

Full Text: PDF(693KB)
Learn to light a candle in the darkest moments of someone’s life. Be the light that helps others see; i

Full Text: PDF(874KB)
You're not going to master the rest of your life in one day. Just relax. Master the day. Than just keep

Full Text: PDF(274.79KB)
Raise your words, not voice. It is rain that grows flowers, not thunder. Rumi

OnlineFirst Full-Text PDF
Seek knowledge from cradle to the grave. Prophet Muhammad (Peace be upon him)

Full Text: PDF(156.81KB)
Don't ruin a good today by thinking about a bad yesterday. Let it go. Anonymous

Full text (pdf, 60 MiB)
Why complain about yesterday, when you can make a better tomorrow by making the most of today? Anon

Full text (pdf, 19 MiB)
Knock, And He'll open the door. Vanish, And He'll make you shine like the sun. Fall, And He'll raise

PDF Full-Text (704 KB)
Pretending to not be afraid is as good as actually not being afraid. David Letterman

Full Text (PDF, 1945 KB)
Almost everything will work again if you unplug it for a few minutes, including you. Anne Lamott

Idea Transcript


Modeling gating charge and voltage changes in response to charge separation in membrane proteins Ilsoo Kima, Suman Chakrabartyb, Peter Brzezinskic, and Arieh Warshela,1 a Department of Chemistry, University of Southern California, Los Angeles, CA 90089; bPhysical and Materials Chemistry Division, National Chemical Laboratory, Pune 411008, India; and cDepartment of Biochemistry and Biophysics, Arrhenius Laboratories for Natural Sciences, Stockholm University, SE-106 91 Stockholm, Sweden

Contributed by Arieh Warshel, June 19, 2014 (sent for review May 5, 2014)

bacterial reaction center proton/electron transfer

| membrane potential | electrogenicity |

T

he use of electrometric techniques to study time-dependent membrane potentials has been exploited in elucidating charge motions in membrane proteins, thereby offering mechanistic insights into the function of membrane-bound channels and pumps (1). For example, data from time-resolved measurements of voltage changes across the redox-driven proton pump cytochrome c oxidase have been used to propose a sequence of electron and proton-transfer events within the enzyme (2). These studies enabled observation of proton-transfer reactions that are typically “invisible” when using spectroscopic techniques and to link these events to electron-transfer reactions. Unfortunately, a correlation of the observed voltage changes with the number and displacements of transferred charges is challenging because the nature of the dielectric response in a protein–membrane system is complex. Thus, it is not straightforward to use macroscopic models to obtain a unique relationship between internal charge motions and the generated potential. The problem is particularly serious if the dielectric or electrostatic response changes significantly in the time range of the experiment. For example, in the case of cytochrome c oxidase a voltage change may be generated by the relaxation of water molecules in an intraprotein proton pathway approximately perpendicular to the membrane surface in response to an internal proton transfer approximately parallel to the membrane surface (see water molecules stretching across the distance from the membrane surface to Glu286 and proton transfer from Glu286 to the catalytic site, respectively; Fig. S1). The voltage change associated with water relaxation would be www.pnas.org/cgi/doi/10.1073/pnas.1411573111

incorrectly interpreted as that of a proton transfer to the heme a3 propionate D, perpendicular to the membrane surface (Fig. S1), which is a key step of the proton-pumping process. In other words, in this case a direct translation of the measured voltage into a charge-transfer distance may not allow for a correct identification of the proton-transfer events. Even though interesting macroscopic studies of such chargetransfer events have been presented (3) it is still a challenge to quantitatively correlate the observed voltage changes to assumed mechanisms. The difficulties arise from both the uncertainties in the macroscopic treatments and the fact that, in contrast to standard cases with a constant applied potential, here the electrode potential is allowed to vary in response to the charge separation. Overcoming the macroscopic uncertainties by applying a fully microscopic study is unfortunately extremely challenging. To resolve the above issues we have used our recently developed coarse-grained (CG) model for the electrode external voltage effect (4), where we introduced a self-consistent treatment that allows the electrode potential to equilibrate with intraprotein charge-transfer reactions. The CG simulations were then validated by reproducing the results from measurements of voltage changes associated with electron and proton transfer in bacterial photosynthetic reaction centers (RCs) (5). 1. Conceptual Considerations Our basic strategy is to consider both experimentally and computationally a well-defined system where the elementary chargeseparation steps are known and to determine the voltage changes and gating charges that arise in response to chargeseparation events. As a start we consider in Fig. 1A a simple schematic system that is set up to monitor a charge-separation event in a membrane. Such an event leads to accumulation of electrolyte charges near the membrane where the opposite charges Significance Determining voltage changes in response to charge separation within membrane proteins offers fundamental information on mechanisms of charge transport and displacement processes. However, correlating the observed voltage changes with the assumed mechanism is extremely challenging owing to uncertainties in macroscopic treatments and the fact that the electrode potential is not fixed. Here, we develop a coarse-grained model that determines the gating charges and the electrode potentials upon charge transfers in membrane proteins. The validity of the model is confirmed by comparing the calculated and measured voltage changes associated with electron and proton transfer in a bacterial photosynthetic reaction center. This model allows one to gain more confidence in the computational analysis than that obtained by continuum electrostatic assumptions. Author contributions: I.K., P.B., and A.W. designed research; I.K., S.C., P.B., and A.W. performed research; I.K., P.B., and A.W. analyzed data; and I.K., P.B., and A.W. wrote the paper. The authors declare no conflict of interest. 1

To whom correspondence should be addressed. Email: [email protected].

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10. 1073/pnas.1411573111/-/DCSupplemental.

PNAS | August 5, 2014 | vol. 111 | no. 31 | 11353–11358

BIOPHYSICS AND COMPUTATIONAL BIOLOGY

Measurements of voltage changes in response to charge separation within membrane proteins can offer fundamental information on mechanisms of charge transport and displacement processes. A recent example is provided by studies of cytochrome c oxidase. However, the interpretation of the observed voltage changes in terms of the number of charge equivalents and transfer distances is far from being trivial or unique. Using continuum approaches to describe the voltage generation may involve significant uncertainties and reliable microscopic simulations are not yet available. Here, we attempt to solve this problem by using a coarse-grained model of membrane proteins, which includes an explicit description of the membrane, the electrolytes, and the electrodes. The model evaluates the gating charges and the electrode potentials (c.f. measured voltage) upon charge transfer within the protein. The accuracy of the model is evaluated by a comparison of measured voltage changes associated with electron and proton transfer in bacterial photosynthetic reaction centers to those calculated using our coarse-grained model. The calculations reproduce the experimental observations and thus indicate that the method is of general use. Interestingly, it is found that charge-separation processes with different spatial directions (but the same distance perpendicular to the membrane) can give similar observed voltage changes, which indicates that caution should be exercised when using simplified interpretation of the relationship between charge displacement and voltage changes.

Fig. 1. (A) A schematic charge development and (B) voltage generation (in millivolts) in the absence of an applied external potential (evaluated by calculations), upon charge separation in the model system where the two charges [ARG(+) on the left and GLU(−) on the right] are located at the membrane boundary, separated by about 40Å. This system is a simplified representation of the RC system without internal charges embedded in the RC protein. The vertical distance of charge separation is comparable to that of 32 Å between P (a bacteriochlorophyll dimer) and QA (quinone) in the RC system (Fig. 4).

calculate Q0 for different ΔV . Unfortunately, in contrast to standard cases with a constant applied potential, here we do not know the final ΔV or Q0. However, we may try to determine ΔV by evaluating the electrolytes distribution for different ΔV and selecting the specific ΔV that leads to the lowest total free energy. More specifically, when we consider a fixed applied potential we obtain a well-defined gating charge, but here our system responds to the charge separation by changing the electrode potential. To analyze such a system we perform CG calculations (section 2) with different assumed electrode potentials and ask which one is at equilibrium with the charge separation by finding the lowest free energy of the whole system (Fig. 2). Although this procedure seems to be the most logical treatment, it might involve some difficulties (discussed below) and will be used here in an exploratory way while considering other options to obtain a stable generated voltage. With the above considerations, we can turn to the experimental study of the voltage generated by the charge separation in bacterial photosynthetic RCs. The relevant analysis has been done previously with the experimental setup that is described in detail in ref. 5. This setup determines the voltage changes associated with the translocation of a charge qda between a donor (d) and an acceptor (a) in a single RC (which is placed with the membrane perpendicular to the field): ΔVda =

are accumulated near the electrodes (we consider a nonequilibrium condition where the total charge is conserved on each side of the membrane). In case of sufficient screening by the electrolytes, the potential between the electrodes is similar to the potential difference across the membrane. Our CG model allows us to evaluate the charge accumulated near the left electrode, which sums up to −Q0 (with an opposite charge near the right electrode). This charge is defined below as the “gating charge,” in analogy with voltage-gated ion channels (4, 6, 7). In addition, we can consider the voltage generation (Fig. 1B). When the internal membrane charge is generated, electrolytes and electrode charges move and the corresponding current ðiÞ satisfies the relation Z ΔV =

Q0 = C

tmax

iðtÞdt C

;

[1]

where C is the capacitance. In addressing the challenge of obtaining Q0, we note that we can use our CG model and

ΔQda qda αda = ; CL CL

[2]

where ΔQda is a displacement charge that flows through the external circuit, CL is the capacitance of the interfacial membrane layer, and αda is a parameter (often called dielectric distance) that was estimated phenomonologically in ref. 5 as αda =

«L dda ; «da DL

[3]

where dda is the distance (perpendicular to the membrane surface) between acceptor and the donor, DL is the thickness of the membrane around the RCs, «da is the dielectric constant in the interior of the RCs in the segment between the donor and acceptor, and «L is the average dielectric constant of the membrane region. The estimate of αda through Eq. 3 (ref. 5) is not certain because the macroscopic assumptions used (8) to determine the parameters that define αda in Eq. 3 are qualitative. Thus, we use the above estimate only as a hint. Fortunately, we are not bound by these assumptions because we are only interested in the value of ΔQda ; which would be determined by our CG approach.

Fig. 2. The procedure of finding the gating charge and the electrode potential for the hypothetical system. The total free energy (red) is evaluated for different voltages using Eq. 7 and the minimum free energy is found at −70 mV. The gating charge (∼0.4 e) at this voltage was estimated using Eq. 6, by considering the difference between the integrated electrolyte distribution (Insets) before (blue) and after (red) the charge separation. Notice that the free energy profile as a function of external (electrode) voltage is quadratic, a feature that appears in any system where the charging (by external voltage in the present study) follows the linear response approximation.

11354 | www.pnas.org/cgi/doi/10.1073/pnas.1411573111

Kim et al.

(9), where the external potential induces surface charges and creates the corresponding displacement vector D0 . In this case we have D0 = σ f

2. Key Features of the CG Modeling Approach Our CG treatment is aimed at modeling the protein–membrane system and its interaction with the external potential. The CG simulation system for the present study has been constructed as described in detail elsewhere (4) and in Supporting Information, Section S3. The simulation system constitutes a simulation box that explicitly includes the membrane, protein, and a grid representing the electrolyte solution. We also add a “bulk region” far away from both the membrane and electrode surfaces, as a way for spanning the space between the membranes and the electrodes without using an enormous grid. The specific treatment of the electrode potential is outlined in ref. 4 and Supporting Information, Section S3. Our CG model solutions consider the electrolytes by using a self-consistent grid-type approach where the residual charges at each grid point represent the charges of the electrolytes ðqgk Þ. The electrostatic potential on each grid point, ϕj (in kilocalories per mole), is expressed as ϕj =   Vext j + 332

X qp k k

«eff jk rjk

+ 332

X qg k ; «wat rjk k>j

where σ f is the surface charge. With the corresponding external potential (Supporting Information, Section S3) and the rest of the CG terms we determine iteratively the membrane potential and the electrolyte charges and also evaluate the free energy of the protein charges in the presence of this potential. An alternative strategy to represent the electrodes is to replace the treatment of Eq. 5 by a finite grid of point charges on the electrode. Both treatments have been described and validated in ref. 4. With the CG model we have the unique ability to evaluate the charge distribution of the electrolytes and the gating charge, rather than its continuum approximation (see discussion in ref. 4). This is done by using the procedure described in the caption of Fig. 3, where we obtained the gating charge by Zz′ Qgate ð≡ ΔQ0 Þ =

  ΔΔqgrid ðV ; ZÞ ΔZ dZ;

[6]

−∞

where ΔΔqgrid is the difference in the accumulative sum of Δqgrid before and after charge translocation and Z′ is the point to the left of Z where the electrolyte charge distribution near the membrane changes sign. At this point, the integrated charge reaches a plateau and then starts to decrease. The overall CG protein/membrane energetics in the presence of electrolytes and the external potential have been refined in the present study by including the penalty for the polarization energy of the system (10). Thus we express the total free energy as: ΔGtot = ΔGCG + 332

X qpj qgk

+ 332

«eff kj rjk X X 1 1 + V ext qp + V ext qg ; 2 j j j 2 j j j k;j

[4]

where rjk (in angstroms) are the distances between corresponding points and Vjext represents the electrode potential on the ith grid point (described below). Here, qpk is the charge of the kth protein residue [these charges are evaluated by a Monte Carlo (MC) procedure (4)] and qgk is the point charge at the kth grid point (representing the excess net charge of the kth volume element). «eff jk = 1 + A½1 − expð−0:5rjk Þ is a distant dependent dielectric constant with an amplitude (A) of either 60 or 80 in the present study (Table 1) and «wat (= 80) is the dielectric constant of the bulk water. The final set of the grid charges fqg g are obtained iteratively (Supporting Information, Section S3). To model the effect of the external potential one can consider formally the membrane/protein/water system as a capacitor. In the present work we use the well-known macroscopic capacitor model



X qgj qgk k>j

«wat rjk

[7]

where ΔGCG is the free energy of the CG protein and membrane system (6), and the half fraction (1/2) reflects the polarization penalty. 3. Results and Discussion To evaluate the model by reliable experimental data with a well-defined system, we measured light-induced voltage changes in photosynthetic RCs from Rhodobacter sphaeroides that were reconstituted in phospholipid vesicles, which were adsorbed to a Teflon film (5). The primary data are summarized in Supporting Information, Section S2 (Figs. S2 and S3). The RCs are membrane-bound protein complexes that convert light energy into electrochemical energy (see Fig. 4 and refs. 11 and 12 for more details on this system). The absorption of light by the donor, P (a bacteriochlorophyll dimer), is followed in time by a series of sequential electron-transfer reactions, which

Table 1. Comparison of measured and calculated electrogenicity Reaction PQAQB → P+QA−QB P+QA−QB → P+QA QB− P+QA− QB− + 2H+ → P+QAQBH2

Normalized measured electrogenicity

Normalized calculated electrogenicity*

Normalized calculated electrogenicity†

Normalized calculated electrogenicity§

1 0 0.14

1 (0.68) 0 (0) 0.38 (0.25)

1 (0.54 e) 0.035 (0.019 e) 0.13 (0.070 e)

1 (1.8 mV) 0.10 (0.18 mV) 0.14 (0.26 mV)

See also Tables S1 and S2. The electrogenicities were evaluated, respectively, using *Eq. 3 with a uniform dielectric constant, †gating charge, and §electrode potential that reflects the CG generated voltage upon charge transfers (Fig. 7). The amplitude (A) of «eff (see text) is 80. However, when A is reduced to 60, the normalized (absolute) electrogenicity in terms of the gating charge is 1(0.210e), 0.10 (0.0210e), and 0.381 (0.081e), respectively, from the first to third reaction. The numbers in parentheses are absolute electrogenicities.

Kim et al.

PNAS | August 5, 2014 | vol. 111 | no. 31 | 11355

BIOPHYSICS AND COMPUTATIONAL BIOLOGY

Fig. 3. The CG model of the system used in the present work: the protein/ membrane system (region I), the electrolyte grid (regions III and IV), and the bulk region (Bulk) between electrolyte grid and the electrodes.

[5]

Fig. 4. The RC system and the key cofactors involved in the charge separation process. The vertical distance (that is normal to membrane) is 25, 0, and 10 Å, respectively, for the charge transfer reactions considered in the present study (Table 1). The membrane width, shown in dashed lines, is 40 Å.

stabilize the charge-separated state for progressively longer times. In the initial step, an electron from P is transferred to a pheophytin, ΦA, and then consecutively to the tightly bound primary quinone acceptor, QA, and the loosely bound secondary acceptor, QB. After absorption of two photons, the doubly reduced quinone binds two protons forming QBH2, which dissociates from the RC and is replaced by an unprotonated quinone from the quinone pool in the membrane. In the present study, the isolated RCs were incorporated into a phospholipid layer, which was adsorbed onto the surface of a Teflon film separating two electrolyte-filled compartments (see refs. 5, 13, and 14 for a description of the system). Illumination of the RC-lipid layer generated voltage changes (Fig. S2) associated with electron and proton-transfer reactions. The increase in voltage induced by the first flash at t = 0 was associated with the charge separation PQA → P+QA− or PQAQB → P+QAQB− in the absence and presence of QB, respectively (the PQA → P+QA− charge separation time constant was 50 quinones per RC) and 20 μM cytochrome c2+. The time between the flashes was 10 s (voltage changes were about the same for time delays 1–20 s). The initial increase in voltage at t = 0 after the first flash is associated with the charge separation PQAQB → P+QA−QB. The small decrease in voltage with a time constant of ∼200 μs is associated with a structural relaxation as a response to the charge separation (see Supporting Information, Section S2; the electron transfer P+QA−QB → P+ QAQB− is not electrogenic). The slower increase in voltage is associated with reduction of P+ by cytochrome c2+ with a time constant of ∼20 ms. The increase in voltage at t = 0 after the second flash is associated with the charge separation PQAQB− → P+QA−QB−, whereas the following increase with a time constant of ∼1.2 ms (see B) is associated with proton uptake to the doubly reduced quinone, P+QA−QB− + 2H+ → P+QAQBH2. The slower increase in voltage is associated with reduction of P+ by cytochrome c2+. (B) The difference between voltage changes after the second and first flashes represents proton uptake to the doubly reduced quinone, QA−QB− + 2H+ → QAQBH2. The smooth line is a fit to a single exponential with a time constant of 1.2 ms. Other conditions: 210 mM KCl, 10 mM CaCl2, 2.5 mM sodium citrate, 2.5 mM Pipes, 2.5 mM HEPPS, 2.5 mM CHES, and 2.5 mM CAPS at pH 8.0; temperature, 22 ± 1 °C.

Kim et al.

state. In principle the next step should involve MC evaluation of the RC ionization state in the presence of the electrolyte charges. This expensive procedure would consider the possible change in the ionization state of the RC charged groups as a result of the equilibration of the electrolytes. However, we felt that handling the possible instability of a fully iterative cycle should be left to subsequent studies. A key element in our treatment is the use of the RC ionization state that corresponds to the PQ initial state. This strategy reflects the assumption that the changes in ionization states (and the corresponding protontransfer processes) are slower than that of the fast charge separation process (see discussion in section 4 and Supporting Information). The calculations were repeated for different sets of applied potentials as described in Fig. 6 and the result with the lowest absolute free energy was taken as the relevant result (note, however, the discussion below). The calculated electrolyte distribution for the PQAQB to P+QA−QB charge separation is presented in Fig. 6, Inset and the resulting gating charge is about 0.54e, which is reasonable, considering the uncertainties in the estimate of ref. 5. It should also be noted that the calculated gating charge was found to be almost independent of the applied potential (Fig. 6), as is the case for the model system of Fig. 2. This finding is important since the calculated dependence of the overall free energy on the electrode potential might still be tentative, because some of the contributions (e.g., the energy of polarizing the electrolytes) might need further refinement. Note, however, that the change between the positions of the minimum for the PQAQB and P+QA−QB states is in the range of 0 with an MC convergence error bar of ∼35 mV. Although the method still needs to be improved (see section 4), the

4. Concluding Remarks This work explored the relationship between charge separation processes in membrane proteins and the corresponding gating charges as well as changes in the electrode potentials (i.e., voltage generation). The consideration of the RC test case gave overall encouraging results and some confidence in our ability to use the CG approach in studies of more challenging problems. The need for CG analysis introduced here has been dictated by major uncertainties in the use of macroscopic models and approaches in studies of membrane–electrolytes–electrode systems. This does not mean that macroscopic models like the one used by Stuchebrukhov and coworkers (3) cannot provide useful information, but we

Fig. 7. The alternative approach of finding the electrode potentials (i.e., voltage generation) for the [P+Q−AQB to P+QAQ−B] process. The voltage generated by charge separation seems independent of the external potential, yielding about 2 mV at two applied different external potentials (A) 200 mV and (B) −200 mV, respectively, as shown in figure.

Kim et al.

PNAS | August 5, 2014 | vol. 111 | no. 31 | 11357

BIOPHYSICS AND COMPUTATIONAL BIOLOGY

Fig. 6. The procedure of determining the gating charge and the electrode potential for the [PQAQB to P+QA−QB] process. The free energy minimum of the reaction is located around +200mV, yielding (using Eq. 6) a gating charge of ∼0.54e, by considering the difference between the integrated electrolyte distribution (Insets) before (blue) and after (red) the charge separation (Inset). The electrode potential (at free energy minima) is in the range 0∼35 mV. The free energy profiles as a function of the external potential were determined using Eq. 7.

result is in the same direction as the observed value reported in ref. 5, where the estimated voltage change is about 1.5 mV (figure 4 in ref. 5). However, whereas we continue to explore the origin of the difficulties to pinpoint the change in potential by calculating the minimum of the free energy/potential correlation we also explored another approach (Fig. 7) for determining voltage changes that was found to be more effective for this specific purpose. That is, we calculated the potential changes near the electrodes for different values of the applied external potential (Fig. 7 and Table 1). Fortunately, the evaluated difference appeared to be insensitive to the applied potential and was found to be about 2 mV, in a good agreement with the observe change (5) as well as the related estimate of the effect of charge separation in cytochrome c oxidase (3). We also explored the gating charge for the [P+ Q− A QB to + P QAQ−B] process and obtained a value close to 0, in agreement with the corresponding observed change (Fig. 8A and Table 1). Next, we examine the gating charge for the [P+Q−AQ−B + 2H+ to P+QA QBH2] process. In this case, we obtained about 0.13e, which is also in a good agreement with the experimentally observed value (Fig. 8B, Table 1, and Table S2). Finally, Fig. 9 shows the relative change in voltage as a function of the charge-transfer distance (Supporting Information, Section S2). As seen in the graph, there is a close to linear dependence (c.f. Eq. 3) for the charge transfer from the surface on the periplasmic side to QA. In trying to analyze the origin of this effect, we found that changing the dielectric constant for the charge–charge interactions affects the protonation states of protein charged residues, which influences the gating charges (Table 1). On the other hand, it was found that a change in the dielectric constant for interactions between the protein residues without changing the protonation states does not lead to any change in the potential. This, however, is not directly related to the dielectric constant in Eq. 3, because the corresponding derivation assumed a macroscopic dielectric for the whole protein and thus may reflect the dielectric for the interaction between the electrolytes and the protein charges (or other elements). Such a treatment also included implicitly the effective dielectric for the response of the protein to the external potential, which is treated more explicitly in the microscopically defined CG model. Thus, we need further studies to elucidate the microscopic origin of the trend of Fig. 9, but we are encouraged by the ability to reproduce this trend.

Fig. 8. Showing the evaluation of the gating charges for (A) the P+QA−QB → P+QA QB− reaction and (B) the P+QA− QB− + 2H+ → P+QAQBH2 reaction in the RC system. The corresponding gating charge (about 0.03e in A and 0.13e in B) was evaluated using Eq. 6, by considering the difference between the integrated electrolyte distribution before (blue) and after (red) the charge separation.

believe that it is essential to use a more microscopic approach and to compare these results to those obtained from experimental studies. Such studies establish the foundation for interpretation of results from studies of more complex systems such as cytochrome c oxidase. In this respect, we find it instructive that the results for the PQAQB → P+QA−QB and PQAQB → P+QAQB− steps (obtained by combining two steps in the table) are very similar although they correspond to charge movement in different directions. This means that the direction of the charge separation cannot be determined in a simple way by just measuring the gating charge (interestingly, the same results are also predicted by Eq. 3). In other words, we find that charge-separation processes with different spatial directions can give similar observed voltage changes. Another case is the proton uptake to QB. Here, the observed (and calculated) voltage change is significantly smaller than that predicted, assuming that the voltage is proportional to distance in the RC. Part of this effect is probably due to the fact that the proton transfer occurs at a high dielectric region. The above observations indicate that caution should be exercised when using simplified relations between charge-displacement distances and observed voltage changes. In other words, the analysis is not simple because it depends on the microscopic nature of the dielectric environment and the dielectric relaxation times and also cannot be determined with certainty using macroscopic formulations. Another point of interest is the finding that the results depend strongly on whether or not we allow for proton reequilibrium after the charge-separation process. Here, we assume that the PQAQB → P+ QA− QB charge separation process is faster than the subsequent proton transfer between different protein ionizable groups or between such groups and bulk solvent (see also Supporting Information). This assumption is justified by the finding that allowing proton equilibration during the first step leads to poor agreement between the calculated and observed voltage changes (Supporting Information, Section S4). However, the possibility of coupled electron–proton transport should be clearly simulated for the slower steps and at present 1. Drachev LA, et al. (1974) Direct measurement of electric current generation by cytochrome oxidase, H+-ATPase and bacteriorhodopsin. Nature 249(455):321–324. 2. Belevich I, Bloch DA, Belevich N, Wikström M, Verkhovsky MI (2007) Exploring the proton pump mechanism of cytochrome c oxidase in real time. Proc Natl Acad Sci USA 104(8):2685–2690. 3. Sugitani R, Medvedev ES, Stuchebrukhov AA (2008) Theoretical and computational analysis of the membrane potential generated by cytochrome c oxidase upon single electron injection into the enzyme. Biochim Biophys Acta 1777(9):1129–1139. 4. Dryga A, Chakrabarty S, Vicatos S, Warshel A (2012) Coarse grained model for exploring voltage dependent ion channels. Biochim Biophys Acta 1818(2):303–317. 5. Brzezinski P, Messinger A, Blatt Y, Gopher A, Kleinfeld D (1998) Charge displacements in interfacial layers containing reaction centers. J Membr Biol 165(3):213–225. 6. Dryga A, Chakrabarty S, Vicatos S, Warshel A (2012) Realistic simulation of the activation of voltage-gated ion channels. Proc Natl Acad Sci USA 109(9):3335–3340. 7. Kim I, Warshel A (2014) Coarse-grained simulations of the gating current in the voltage-activated Kv1.2 channel. Proc Natl Acad Sci USA 111(6):2128–2133. 8. Läuger P, et al. (1981) Relaxation studies of ion transport systems in lipid bilayer membranes. Q Rev Biophys 14(4):513–598. 9. Jackson JD (1998) Classical Electrodynamics (Wiley, New York), 3rd Ed.

11358 | www.pnas.org/cgi/doi/10.1073/pnas.1411573111

Fig. 9. The relation between the measured voltage (per unit charge) and the position within the membrane (i.e., for the different studied reactions). SP and SC are the RC surfaces on the periplasmic and cytoplasmic sides, respectively. See Supporting Information, Section S2 for details.

the most effective way for exploring this problem is the use of our time-dependent CG model for proton transport (16). Fortunately we obtained for the PQA−QB → P+QAQB− similar results with and without proton reequilibration (Table S3). In summary, the present CG study provides a powerful tool for interpretation of data from studies of transmembrane voltage changes and for correlating the observed electrogenicities with the corresponding molecular events. This is particularly significant because in almost all studies it is assumed that the measured voltage is proportional to the number of charges and the distance independently of the region where the charge separation occurs within the protein. Our calculations show that this assumption is not valid and thus a modeling approach that considers explicitly the entire system is important. 5. Methods Our general strategy involves a refinement of our recent CG model and the extension of this model to the incorporation of an external potential in the simulation of protein/membrane system. The protein system is treated by a CG model that describes the main chains by an explicit model that represents the side chains as a simplified united atom model, whereas the membrane is described by a grid of nonpolar groups. This CG model provides a more advanced treatment of electrostatic effects than most current CG models (for more details see ref. 4 and Supporting Information, Section S2). ACKNOWLEDGMENTS. We thank the University of Southern California’s High-Performance Computing and Communications Center for computer time. The experimental data with the bacterial reaction centers were obtained in the laboratory of George Feher and Melvin Y. Okamura at the University of California, San Diego. This work was supported by National Institutes of Health Grant GM40283, National Science Foundation Grant MCB-0836400, the Knut and Alice Wallenberg Foundation, and the Swedish Research Council.

10. Warshel A, Levitt M (1976) Theoretical studies of enzymic reactions: Dielectric, electrostatic and steric stabilization of the carbonium ion in the reaction of lysozyme. J Mol Biol 103(2):227–249. 11. Warshel A, Parson WW (2001) Dynamics of biochemical and biophysical reactions: Insight from computer simulations. Q Rev Biophys 34(4):563–679. 12. Okamura MY, Paddock ML, Graige MS, Feher G (2000) Proton and electron transfer in bacterial reaction centers. Biochim Biophys Acta 1458(1):148–163. 13. Brzezinski P, Paddock ML, Okamura MY, Feher G (1997) Light-induced electrogenic events associated with proton uptake upon forming QB- in bacterial wild-type and mutant reaction centers. Biochim Biophys Acta 1321(2):149–156. 14. Brzezinski P, Okamura MY, Feher G (1992) Structural changes following the formation of D+ Q A-in bacterial reaction centers: Measurement of light-induced electrogenic events in RCs incorporated in a phospholipid monolayer. The Photosynthetic Bacterial Reaction Center II, eds Breton J, Vermeglio A (Springer, Berlin), pp 321–330. 15. Drachev L, et al. (1990) Electrogenesis associated with proton transfer in the reaction center protein of the purple bacterium Rhodobacter sphaeroides. FEBS Lett 259(2):324–326. 16. Messer BM, et al. (2010) Multiscale simulations of protein landscapes: using coarsegrained models as reference potentials to full explicit models. Proteins 78(5): 1212–1227.

Kim et al.

Smile Life

When life gives you a hundred reasons to cry, show life that you have a thousand reasons to smile

Get in touch

© Copyright 2015 - 2024 PDFFOX.COM - All rights reserved.