Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2015 Feb 2.
Published in final edited form as: Phys Biol. 2013 Aug 2;10(4):045001. doi: 10.1088/1478-3975/10/4/045001

Folding free energy surfaces of three small proteins under crowding: validation of the postprocessing method by direct simulation

Sanbo Qin 1, Jeetain Mittal 2, Huan-Xiang Zhou 1,*
PMCID: PMC3870194  NIHMSID: NIHMS514182  PMID: 23912849

Abstract

We have developed a “postprocessing” method for modeling biochemical processes such as protein folding under crowded conditions. In contrast to the direct simulation approach, in which the protein undergoing folding is simulated along with crowders, the postprocessing method requires only the folding simulation without crowders. The influence of the crowders is then obtained by taking conformations from the crowder-free simulation and calculating the free energies of transferring to the crowders. This postprocessing yields the folding free energy surface of protein under crowding. Here the postprocessing results for the folding of three small proteins under “repulsive” crowding are validated by those obtained previously by the direct simulation approach. This validation confirms the accuracy of the postprocessing approach and highlights its distinct advantages in modeling biochemical processes under cell-like crowded conditions, such as enabling an atomistic representation of the test proteins.

Keywords: protein folding, macromolecular crowding, cell-like conditions, postprocessing

1. Introduction

It is now widely recognized that the crowded conditions found in cellular environments can significantly affect the thermodynamic and kinetic properties of biochemical processes such as protein folding [1]. Experimental [2-6] and computational studies [7-9] have begun to move beyond a qualitative understanding [10-12] of crowding effects and are aiming to probe quantitative questions regarding the magnitudes of such effects and the nature of the interactions between test proteins and crowders. We have developed a computational approach called postprocessing that opens the door to making quantitative predictions on crowding effects [8]. In contrast to the direct simulation approach [7, 9], in which one simulates the folding process in the presence of crowders and has to resort to a coarse-grained representation of the test protein, the postprocessing approach requires only the simulation in the absence of crowders and thereby enables an atomistic representation of the test protein. Here we validate the postprocessing approach by showing that the folding free energy surfaces of three small proteins under “repulsive” crowding, obtained previously by Mittal and Best (MB hereafter) [9] using direct simulation, are accurately predicted by postprocessing the crowder-free simulation.

In the postprocessing approach, one captures the effects of crowding by calculating how much the free energy surface of a biochemical process such as folding is changed by the crowders [1, 8, 13]. This change is simply the free energy of transferring the test protein in a given state (or microstate) from a dilute solution (modeled by a crowder-free simulation) to a crowded solution. The transfer free energy, Δμ, is given by

exp(ΔμkBT)=exp[Uint(X,R,Ω)kBT]0;crowd (1)

where Uint(X, R, Ω) is the effective interaction energy between the test protein and the crowders when the former has conformation X, position R, and orientation Ω; kB is Boltzmann's constant; T is the absolute temperature; 0;crowd means averaging over the conformation, position, and orientation of the test protein and over the configuration of the crowders; and the subscript “0” signifies that the protein conformations are those sampled in the absence of crowders. For each protein conformation, the averaging over protein position and orientation and crowder configuration can be carried out first, leading to

exp(ΔμkBT)=exp[Gc(X)kBT]0 (2)

Gc(X) can be viewed as the transfer free energy for the protein in conformation X. The postprocessing calculation is exact if the conformational space of the protein is exhaustively sampled so as to cover the important conformational region in the presence of crowders.

In a previous study [14], we have shown that the effects of crowding on the flap open-to-closed population ratio of the HIV-1 protease dimer predicted by the postprocessing approach from the crowder-free simulation agree well with those obtained from direct simulations in the presence of crowders [15]. Of course the conformational difference between the flap open and flap closed states is somewhat limited. Here we present a much more stringent validation of the postprocessing approach. We show that the entire folding free energy surfaces of three small proteins (figure 1), encompassing the fully unfolded state and the fully folded state, obtained by MB in direct simulations with crowders, are well predicted by postprocessing the crowder-free simulation. The validation also highlights the distinct advantages of the postprocessing approach, including superior statistics and the ability to more realistically model the test proteins, the crowders, and their interactions.

Figure 1.

Figure 1

The three proteins studied. Their structures and four different-sized crowders are shown to scale.

2. Computational details

2.1 Calculation of F(Q) from replica-exchange umbrella sampling

MB obtained the folding free energy function, F(Q), along the fraction, Q, of native contacts from replica-exchange simulations with umbrella sampling. F(Q) was calculated by the weighted histogram analysis method (WHAM) [16]. Below we extend WHAM to account for the effects of crowding. In preparation, we briefly summarize the original WHAM algorithm for a set of simulations (referred to by index i) applying biasing potentials Vi(Q) that constrain the protein to different windows along Q. First consider the case where all the simulations are carried out at the same temperature T; for notational simplicity, we will now denote 1/kBT as β. Suppose that ni snapshots (referred to by index j) are sampled from simulation i, and from which a histogram pi(Ql) in different bins, denoted by Ql, are obtained:

pi(Ql)=j=1niH(QijQl)ni (3)

where Qij is the value of Q in snapshot j of simulation i, and H(Qij|Ql) is 1 of Qij is in bin Ql and 0 otherwise. By removing the biasing potential Vi(Q) and introducing an appropriate shift Ai, one obtains the equilibrium probability density:

P(Ql)=pi(Ql)exp{β[Vi(Ql)Ai]} (4)

Better statistics are achieved when the data from different simulations are combined. To that end one rearranges equation (4) and then sums over i, leading to

P(Ql)=ij=1niH(QijQl)iniexp{β[Vi(Ql)Ai]} (5)

The shift Ai can be obtained from equation (4) by using the condition that the sum of pi(Ql) over Ql is 1 (see equation (3)):

exp(βAi)=QlP(Ql)exp[βVi(Ql)] (6)

An implementation of the WHAM involves iterating equations (5) and (6) to convergence. Finally the free energy function is given by

exp[βF(Q)]=P(Q) (7)

Now consider the case where each simulation is at a different temperature Ti (correspondingly 1/kBTi = βi). The goal again is to calculate the equilibrium probability density P(Ql) at temperature T. The difference of the simulation temperature Ti from the desired temperature T amounts to a biasing factor exp[–(ββi)Uij], where Uij is the energy of the protein in snapshot j of simulation i [17]. Defining

mi=j=1niexp[(ββi)Uij] (8)

and accounting for the biasing factor, one finds the counterpart of pi(Ql) as

qi(Ql)=j=1niH(QijQl)exp[(ββi)Uij]mi (9)

and now

P(Ql)=qi(Ql)exp{βi[Vi(Ql)Ai]} (10)

Equations (5) and (6) become

P(Ql)=ij=1niH(QijQl)exp[(ββi)Uij]imiexp{βi[Vi(Ql)Ai]} (11)
exp(βiAi)=QlP(Ql)exp[βiVi(Ql)] (12)

Replica exchange simulations generate equilibrium distributions at multiple temperatures. The snapshots sampled from such simulations are equivalent to those in the case just considered. Iteration of equations (11) and (12) thus yields the equilibrium probability density and hence the free energy function at the desired temperature. MB carried out the simulations both in the absence and in the presence of crowders. The difference between the resulting free energy functions, F(Q) and Fc(Q), indicates the effects of crowding.

2.2 Calculation of F c (Q) by postprocessing

The goal of postprocessing is to generate the equilibrium probability density Pc(Q) and hence the free energy function Fc(Q) under crowding from the crowder-free simulations. For each crowder-free simulation, introducing an additional biasing factor exo[–βGc(Xij)] for each sampled conformation Xij will result in the histogram under crowding. Defining

vi=j=1niexp[(ββi)Uij]exp[βGc(Xij)] (13)

we find the histogram under crowding as

ρi(Ql)=j=1niH(QijQl)exp[(ββi)Uij]exp[βGc(Xij)]vi (14)

Correspondingly the equilibrium probability density under crowding is

Pc(Ql)=ρi(Ql)exp{βi[Vi(Ql)Ai]} (15)

In analogy to equations (11) and (12) we now have

Pc(Ql)ij=1niH(QijQl)exp[(ββi)Uij]exp[βGc(Xij)]iviexp{βi[Vi(Ql)Ai]} (16)
exp(βiAi)=QlPc(Ql)exp[βiVi(Ql)] (17)

2.3 Calculation of exp[–βGc(Xij)]

In MB's simulations, the test protein had purely repulsive interactions with spherical crowders. If we further approximate this interaction energy as infinite when the test protein clashes with a crowder and 0 when no clash occurs with any crowder, then the Boltzmann factor exp[–Uint(X, R, Ω)/kBT] is either 0 or 1. Correspondingly exp[–Gc(X)kB/kBT] is the fraction of clash-free, or allowed attempts to randomly insert the test protein with conformation X into the crowders. We designed an algorithm for calculating the allowed fraction of insertion by mimicking the fictitious insertion process [8]. Subsequently we developed a semi-analytical method, referred to as the generalized fundamental measure theory (GFMT) [18], to substantially speed up the calculation. The GFMT was crucial for the present study since postprocessing calculations on a large number of protein conformations were required.

For a test protein in conformation X, analogous to the original fundamental measure theory [19, 20], the GFMT predicts the transfer free energy Gc(X) as

Gc(X)=Πcvp+γcsp+κclpkBTln(1ϕ) (18)

where ϕ is the volume fraction of the crowders; Πc, γc, and κc are the osmotic pressure, surface tension, and bending rigidity, respectively, of the crowder suspension; and vp, sp, and lp are the volume, surface area, and linear size, respectively, of the test protein. Equation (19) equates the transfer free energy to the work required to create a cavity in the crowder suspension to accommodate a test particle. The first term on the right-hand side is the work for creating the cavity volume, the second term is work for creating the interface area between the cavity and the crowder suspension, and the third term accounts for the curvature of this interface. The original fundamental measure theory [19, 20] was restricted to hard convex test particles. Our generalization allows atomistic test proteins to be treated.

The generalization involved appropriately defining the geometric quantities vp, sp, and lp for an atomistic test protein. The calculation of vp and sp was based on the crowder-excluded surface, which depends on the crowder radius. This surface, when the probe radius is 1.4 Å (modeling a water molecule), is more commonly referred to as the molecular surface. The calculation of lp involved a surface that is slightly inflated from the crowder-excluded surface. This inflated surface is composed of all points where rays emanating from the center of the protein first hit the closest approaching crowders around the protein; lp was calculated as the radius of gyration for these hit points.

Πc, γc, and κc are crowder-only quantities. Suppose that the crowder suspension consists of multiple species, with number density cα, linear size lα, surface area sα, and volume vα for species α. Note that

ϕ=αcαvα (19)

Let the total number density, the number-averaged linear size and surface area of the crowders, be c, c and c, respectively:

c=αcα (20)
lc=αcαlαc (21)
sc=αcαsαc (22)

Then

βΠc=Y+lcscY2+sc3Y312π (23)
βγc=lcY+sc2Y28π (24)
βκc=scY (25)

where Y = c/(1 – ϕ).

The GFMT implemented for proteins represented at the all-atom level closely reproduced results obtained by the insertion algorithm for the effects of crowding on folding and binding free energies [18]. Here (as done previously [14]), for the purpose of validating the postprocessing approach against direct simulations, we kept MB's coarse-grained representation of one bead per residue for the three test proteins. In this representation, proposed by Karanicolas and Brooks [21], the beads are centered at the Cα atoms and have radii of 4.0 ± 0.8 Å; the interactions between residues in contact in the native structure are modeled as attractive and any other residue-residue contacts formed during conformational sampling are repulsive. MB found that this representation resulted in two-state folding behavior.

The postprocessing predictions involved calculating vp, sp, and lp at a given crowder radius for each protein conformation sampled in a crowder-free simulation. Our computer code for generating the crowder-excluded surface (to calculate vp and sp) and inflated surface (to calculate lp) was applied previously to proteins represented as all atoms with van der Waals radii [18]. This code was used here for the three proteins represented as Cα beads with enlarged radii.

3. Results

The three small proteins studied by MB are prb, protein G, and TNfn3, with 47, 56, and 90 residues respectively (figure 1). The 100 × 100 × 100 Å3 simulation box contained a test protein without or with various numbers (Nc) of spherical crowders (radius: Rc). The interactions between the Cα beads of the protein and the crowders were purely repulsive. The nominal volume fraction, ϕ0, of the crowders equals 4πNcRc3(3×106). Below we compare their simulation results for Fc(Q) against those predicted by the postprocessing approach, which uses only the simulations without crowders. We also compare the resulting effects of crowding on the folding free energy and on the folding rate constant.

3.1 Folding free energy function under crowding

We first made the postprocessing calculations for Fc(Q) by assuming ϕ0 as the actual crowder volume fraction ϕ. The results generally show reasonable agreement with MB's direct simulation results, but as illustrated in figure 2a for protein G in the presence of 8 Å crowders, there is a systematic underestimation of crowding effects.

Figure 2.

Figure 2

Comparison of Fc(Q) results from MB's direct simulations and predicted by the postprocessing approach. Circles, diamonds, and triangles show the simulation results for protein G in the presence of 8 Å crowders at ϕ0 = 0.05, 0.15, and 0.26, respectively. The dashed curve shows the free energy function in the absence of crowders; the three solid curves show Fc(Q) for the three ϕ0 values by postprocessing the crowder-free simulations. The temperature is 310 K. The postprocessing results are obtained by either (a) assuming ϕ0 as the actual crowder volume fraction ϕ; or correcting for the finite size of each simulation system, through (b) carving out the region in and around the real test protein inaccessible to a crowder in calculating ϕ or (c) treating the insertion of a fictitious test protein into the crowders around the real test protein as insertion into the mixture of the real test protein and the crowders.

We then realized that the simulation system has a small, finite size. The test protein presents a volume that is inaccessible to the crowders; the actual volume fraction of the crowders around the test protein is thus higher than that (i.e., ϕ0) calculated by assuming the absence of the test protein. We thus re-calculated ϕ by carving out from the total volume the region in and around the test protein inaccessible to each crowder (figure 3). The inaccessible volume vinac, could be calculated by rolling around the protein a spherical probe with the crowder radius Rc, or could be estimated by Steiner's formula [22]

vinac=vp+Rcsp+lpsc+vc (26)

where sc=4πRc2 and vc=4πRc33, respectively, are the surface area and volume of the spherical probe. We found that these two ways gave very similar results for vinac and all the data presented below are by using equation (26). The correction in ϕ brings the postprocessing predictions into close agreement with MB's direct simulation results for Fc(Q), as illustrated in figure 2b for protein G interacting with 8 Å crowders.

Figure 3.

Figure 3

Correcting for the finite size of MB's simulation system. Because the test protein in the direct simulation presents a volume (vinac) inaccessible to the crowders, the actual volume fraction of the crowders is higher than the nominal volume fraction ϕ0. Inserting a fictitious test protein into the crowders around the real test protein is equivalent to inserting into the mixture of the real test protein and the crowders.

Note that calculating the transfer free energy Gc(X) required for predicting Fc(Q) amounts to inserting a fictitious test protein into the crowders around the real test protein (figure 3). The above correction for the finite size of the simulation system by carving out the region in and around the real test protein inaccessible to the crowders could result in overestimation of the crowding effect, since, when inserting the fictitious test protein, the inaccessible volume around the real test protein and the inaccessible volume around a crowder can overlap. This overlap volume is effectively counted twice in the above correction, once in calculating ϕ, and a second time when testing whether an insertion attempt is allowed. The overlap volume increases with increasing sizes of the test protein and the crowders. The resulting overestimation of the crowding effect was significant only for protein G interacting with 16 Å crowders and for TNfn3 interacting with 16 Å or 20 Å crowders.

A better correction for the finite size of the simulation system is to treat the insertion of a fictitious test protein into the crowders around the real test protein as insertion into the mixture of the real test protein and the crowders (figure 3). With this treatment, the postprocessing predictions for Fc(Q) are in good agreement with MB's direct simulation results for all the three proteins interacting with all the different sized crowders. The results for protein G interacting with 8 Å crowders shown in figure 2c are representative of the extent of the agreement.

3.2 Effect of crowding on folding stability

As figure 2 shows, the free energy functions and F(Q) and Fc(Q) define two basins, one for the unfolded state (with free energy minimum at QQU~0.1), and the other for the native state (with free energy minimum at QQN ~ 0.9). The folding stability is measured by the free energy difference, ΔFN–U, which can be calculated from the free energy functions F(Q) as

ΔFNU=kBTln{0Qexp[βF(Q)]dQQ1exp[βF(Q)]dQ} (27)

where Q, with a value ~ 0.5, denotes the location of the free energy maximum separating the two stable states. An analogous free energy difference can be calculated for the situation under crowding. In figure 4 we compare the postprocessing predictions and MB's direct simulation results for the change in the free energy difference by crowding, ΔΔN–U, for the three proteins under various levels of crowding. It can be seen that, after correcting for the finite size of the simulation system, there is good agreement between postprocessing and direct simulation.

Figure 4.

Figure 4

Comparison of the postprocessing predictions and MB's direct simulation results for the effects of crowding on the folding stability of the three proteins. Circles, diamonds, triangles, and squares show the simulation results for crowders with radii at 8, 12, 16, and 20 Å, respectively. The corresponding curves show the postprocessing predictions after correcting for the finite size of each simulation system, by either (a-c) the mixture treatment or (d) the carving out treatment. (a) prb at 320 K; (b) and (d) protein G at 320 K; (c) TNfn3 at 300 K.

Some of the direct simulation results are apparently noisy, such as those for TNfn3 in the presence of 12 Å crowders at five ϕ0 values. The noisy data can be attributed to the small number of crowders included in the simulations. For TNfn3 in the presence of 20 Å crowders, there were only 4 to 10 crowders in the simulation box. In contrast, in the postprocessing approach, the test protein fictitiously probes an essentially infinite suspension of crowders. The resulting better statistics yield the smooth dependences of ΔΔFN–U on ϕ0 shown in figure 4.

3.3 Effect of crowding on folding kinetics

The rate constant for the transition from the unfolded state to the native state can be calculated according to Kramers’ theory [23]:

kf=10Qexp[βF(Q)]dQQUQN[D(Q)]1exp[βF(Q)]dQ (28)

where D(Q) is the effective diffusion coefficient along Q. The folding rate constant under crowding, kfc, can be similarly calculated. If the crowding does not affect D(Q), the we have

kfckf=0Qexp[βF(Q)]dQQUQNexp[βF(Q)]dQ0Qexp[βFc(Q)]dQQUQNexp[βFc(Q)]dQ (29)

MB found that equation (29) predicts accurately the effect of crowding on the folding rate constant, supporting that crowding does not affect D(Q) for the present systems.

In figure 5 we compare the kfc/kf results calculated from equation (29), using either the Fc(Q) predicted by postprocessing or the Fc(Q) from MB's direct simulation, for prb under various levels of crowding. This comparison again validates the postprocessing approach.

Figure 5.

Figure 5

Comparison of the postprocessing predictions and MB's direct simulation results for the effect of crowding on the folding rate constant of prb. Symbols have the same meaning as in figure 4a.

4. Discussion

We have shown that the postprocessing approach, using only the simulations in the absence of crowders, can predict well the entire free energy surfaces of three small proteins under repulsive crowding. Consequently the effects of crowding on the folding stability and the folding rate constant can be well predicted. Although we have previously validated the postprocessing approach against direct simulation with crowders for the flap open-to-closed population ratio [14], there the conformational space involved was relatively limited. In contrast, here the conformational space sampled encompasses the fully folded and fully folded state, and thus presents a much more stringent validation of the postprocessing approach.

To sample the entire folding free energy surface under crowding, MB used a relatively small 100 × 100 × 100 Å3 simulation box. For the largest sized (20 Å) crowders, only a few (4 to 10) of them were in the simulation box to represent the crowded environment. We found that the finite size of the simulation system has a significant effect. The crowded environment in the direct simulation is different from that if the single test protein were taken away. We found a way to correct for the finite size effect, by treating the insertion of a fictitious test protein into the crowded environment in the direct simulation as insertion into the mixture of the real test protein and the crowders. In our previous validation study [14], the simulation was much larger (at 510 × 510 × 510 Å3) and hence the finite size effect would have been negligible. We emphasize that the correction for the finite size effect was necessary only because we wanted to compare the postprocessing predictions against the direct simulation results. In all our other applications [8, 18, 24-28], we use an essentially infinite crowder suspension for the insertion to calculation the transfer free energy.

In addition to avoiding the finite size effect, in the postprocessing approach the crowder configurations are exhaustively sampled because the fictitious test protein is inserted everywhere in the essentially infinite crowder suspension. The postprocessing results thus have better statistics than the direct simulation counterparts.

The postprocessing calculation is formally exact if the conformational space of the protein is exhaustively sampled by crowder-free simulations, so as to cover the important conformational region in the presence of crowders. A potential concern is whether crowder-free simulations in practice can achieve this goal. The present validation study as well as the previous one [14] allays this concern. However, we do note that in both studies the test proteins had purely repulsive interactions with the crowders. It is possible that attractive protein-crowder interactions may alter the energy landscape of a protein such that new low-energy regions emerge in the presence of crowders. These regions could be under-sampled in crowder-free simulations, leading to errors in the postprocessing calculation. Only further validation will tell the level of these errors.

The most important advantage of the postprocessing approach is the ability to represent the test protein at the atomistic level and its interaction with the crowders more realistically. As observed recently by Gruebele and co-workers [6], the atomistic level representation is required to capture the subtle effects of crowding on protein folding. With the present validation study, a major potential concern regarding the validity of the postprocessing approach is removed at least for repulsive crowders. We can now apply this powerful approach with confidence for realistic modeling of crowding effects.

Acknowledgements

This study was supported by grant GM88187 from the National Institutes of Health.

References

  • 1.Zhou HX, Rivas G, Minton AP. Macromolecular crowding and confinement: biochemical, biophysical, and potential physiological consequences. Annu Rev Biophys. 2008;37:375–97. doi: 10.1146/annurev.biophys.37.032807.125817. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Ai X, Zhou Z, Bai Y, Choy WY. 15N NMR spin relaxation dispersion study of the molecular crowding effects on protein folding under native conditions. J Am Chem Soc. 2006;128:3916–7. doi: 10.1021/ja057832n. [DOI] [PubMed] [Google Scholar]
  • 3.Batra J, Xu K, Zhou H-X. Nonaddtive effects of mixed crowding on protein stability. Proteins. 2009;77:133–8. doi: 10.1002/prot.22425. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Mukherjee S, Waegele MM, Chowdhury P, Guo L, Gai F. Effect of macromolecular crowding on protein folding dynamics at the secondary structure level. J Mol Biol. 2009;393:227–36. doi: 10.1016/j.jmb.2009.08.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Miklos AC, Sarkar M, Wang Y, Pielak GJ. Protein crowding tunes protein stability. J Am Chem Soc. 2011;133:7116–20. doi: 10.1021/ja200067p. [DOI] [PubMed] [Google Scholar]
  • 6.Guo M, Xu Y, Gruebele M. Temperature dependence of protein folding kinetics in living cells. Proc Natl Acad Sci U S A. 2012;109:17863–7. doi: 10.1073/pnas.1201797109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Cheung MS, Klimov D, Thirumalai D. Molecular crowding enhances native state stability and refolding rates of globular proteins. Proc Natl Acad Sci U S A. 2005;102:4753–8. doi: 10.1073/pnas.0409630102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Qin S, Zhou H-X. Atomistic modeling of macromolecular crowding predicts modest increases in protein folding and binding stability. Biophys J. 2009;97:12–9. doi: 10.1016/j.bpj.2009.03.066. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Mittal J, Best RB. Dependence of protein folding stability and dynamics on the density and composition of macromolecular crowders. Biophys J. 2010;98:315–20. doi: 10.1016/j.bpj.2009.10.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Minton AP. Effect of a concentrated “inert” macromolecular cosolute on the stability of a globular protein with respect to denaturation by heat and by chaotropes: a statistical-thermodynamic model. Biophys J. 2000;78:101–9. doi: 10.1016/S0006-3495(00)76576-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Zhou H-X. Protein folding and binding in confined spaces and in crowded solutions. J Mol Recognit. 2004;17:368–75. doi: 10.1002/jmr.711. [DOI] [PubMed] [Google Scholar]
  • 12.Minton AP. Models for excluded volume interaction between an unfolded protein and rigid macromolecular cosolutes: macromolecular crowding and protein stability revisited. Biophys J. 2005;88:971–85. doi: 10.1529/biophysj.104.050351. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.McGuffee SR, Elcock AH. Diffusion, crowding & protein stability in a dynamic molecular model of the bacterial cytoplasm. PLoS Comput Biol. 2010;6:e1000694. doi: 10.1371/journal.pcbi.1000694. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Qin S, Minh DD, McCammon JA, Zhou HX. Method to predict crowding effects by postprocessing molecular dynamics trajectories: application to the flap dynamics of HIV-1 protease. J Phys Chem Lett. 2010;1:107–10. doi: 10.1021/jz900023w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Minh DD, Chang CE, Trylska J, Tozzini V, McCammon JA. The influence of macromolecular crowding on HIV-1 protease internal dynamics. J Am Chem Soc. 2006;128:6006–7. doi: 10.1021/ja060483s. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Kumar S, Bouzida D, Swendsen RH, Kollman PA, Rosenberg JM. The weighted histogram analysis method for free-energy calculations on biomolecules. I. The method. J Comput Chem. 1992;13:1011–21. [Google Scholar]
  • 17.Gallicchio E, Andrec M, Felts AK, Levy RM. Temperature weighted histogram analysis method, replica exchange, and transition paths. Journal of Physical Chemistry B. 2005;109:6722–31. doi: 10.1021/jp045294f. [DOI] [PubMed] [Google Scholar]
  • 18.Qin S, Zhou HX. Generalized fundamental measure theory for atomistic modeling of macromolecular crowding. Phys Rev E. 2010;81:031919. doi: 10.1103/PhysRevE.81.031919. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Rosenfeld Y. Free-energy model for the inhomogeneous hard-sphere fluid mixture and density-functional theory of freezing. Phys Rev Lett. 1989;63:980–3. doi: 10.1103/PhysRevLett.63.980. [DOI] [PubMed] [Google Scholar]
  • 20.Oversteegen SM, Roth R. General methods for free-volume theory. J Chem Phys. 2005;122:214502. doi: 10.1063/1.1908765. [DOI] [PubMed] [Google Scholar]
  • 21.Karanicolas J, Brooks CL., 3rd The origins of asymmetry in the folding transition states of protein L and protein G. Protein Sci. 2002;11:2351–61. doi: 10.1110/ps.0205402. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Steiner J. Über parallele Flächen. Monatsber preuss Akad Wiss. 1840;2:114–8. [Google Scholar]
  • 23.Kramers HA. Brownian motion in a field of force and the diffusion model of chemical reactions. Physica. 1940;7:284–304. [Google Scholar]
  • 24.Batra J, Xu K, Qin S, Zhou H-X. Effect of macromolecular crowding on protein binding stability: modest stabilization and significant biological consequences. Biophys J. 2009;97:906–11. doi: 10.1016/j.bpj.2009.05.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Dong H, Qin S, Zhou HX. Effects of macromolecular crowding on protein conformational changes. PLoS Comput Biol. 2010;6:e1000833. doi: 10.1371/journal.pcbi.1000833. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Tjong H, Zhou HX. The folding transition-state ensemble of a four-helix bundle protein: helix propensity as a determinant and macromolecular crowding as a probe. Biophys J. 2010;98:2273–80. doi: 10.1016/j.bpj.2010.01.052. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Qin S, Cai L, Zhou HX. A method for computing association rate constants of atomistically represented proteins under macromolecular crowding. Phys Biol. 2012;9:066008. doi: 10.1088/1478-3975/9/6/066008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Phillip Y, Harel M, Khait R, Qin S, Zhou HX, Schreiber G. Contrasting factors on the kinetic path to protein complex formation diminish the effects of crowding agents. Biophys J. 2012;103:1011–9. doi: 10.1016/j.bpj.2012.08.009. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES