Important roles of hydrophobic interactions in folding and charge interactions in misfolding of α-helix bundle protein

Qiang Shao*
Drug Discovery and Design Center, Shanghai Institute of Materia Medica, Chinese Academy of Sciences, 555 Zuchongzhi Road, Shanghai, 201203, China. E-mail: qshao@mail.shcnc.ac.cn

Received 11th November 2014 , Accepted 8th December 2014

First published on 8th December 2014


Abstract

Integrated-tempering-sampling molecular dynamics simulation is utilized to investigate the folding of a 67-residue three-α-helix bundle, α3W. Reversible folding and unfolding can be observed in individual trajectories and a total of 28 folding events are achieved within 7 μs simulation, giving sufficient sampling on the configuration space of protein folding. The native-like state with a left-handed topology constitutes the largest fraction of the conformational ensemble sampled by the simulation. In addition, a misfolded state with mirror-image (right-handed) topology is observed with smaller population. The free energy landscape analysis demonstrates that the folding of α3W is initiated by the formation of α-helical secondary structures and is followed by the assembling of folded α-helices to construct tertiary structure. The “correct” α-helix assembling which leads to the native structure is mainly dominated by inter-helical hydrophobic interactions whereas the “incorrect” assembling which leads to a misfolded mirror-image structure is highly affected by not only hydrophobic but also charge interactions. It is speculated on the basis of the present study on α3W and other studies on the B domain of protein A and α3D that the misfolding probability of α-helix bundle proteins is dependent on the strength of intra-protein hydrophobic and charge interactions: proteins containing stronger hydrophobic interactions but weaker charge interactions should have smaller misfolding probability. The importance of intra-protein hydrophobic interactions in preventing protein misfolding has been also seen in our previous studies on β-hairpins. Therefore, the present study along with our previous studies provide comprehensive, atomic-level picture of the folding/misfolding of α-helix bundle and β-hairpin proteins.


1. Introduction

The conformational specificity of a protein reflects its tendency to fold into a unique three-dimensional structure (or set of closely related structures). The failure to fold into functionally correlated native structure might result in severe physiological consequences. For instance, neurodegenerative disorders such as Parkinson's and Alzheimer's diseases are induced by the misfolding of a protein to a β-amyloid like structure which subsequently aggregates to form extracellular amyloid fibrils. Intriguingly, recent multiple molecular simulations have indicated that certain proteins/peptides can adopt alternative conformations in aqueous solution of which the most populated and thus experimentally detectable state is treated as native (folded) state whereas the less populated and experimentally undetectable one can be considered as misfolded state.1–9 Unlike the irreversible protein misfolding and thus the incompatibility of folded and misfolded states of amyloid peptides, the folded and misfolded structures of these proteins with low conformational specificity can coexist in aqueous solution. For instance, multiple β-structured polypeptides (e.g., chignolin, the C-terminal β-hairpin fragment (residues 41–56) from the B1 domain of protein G (GB1p) and its mutants) have been observed to adopt a less populated “out-of-register” β-hairpin structure in addition to the experimentally measured native symmetric hairpin structure.1,2,4–6 In comparison to the folded hairpin structure, the misfolded one has shorter turn region and oppositely orientated cross-strand hydrophobic side chains which are supposed to pack in same direction to form hydrophobic core cluster like in folded structure.

As popular models for in vitro investigation of protein folding, three-α-helix bundles are also geometrically interesting and has potential of misfolding. After the first two helices are aligned, the third one can be packed against either side of the plane involving the former two helices, corresponding to left-handed or right-handed conformation.9 The representative three-α-helix bundle, the B domain of protein A from Staphylococcus aureus (BdpA, 46 residues), has been indicated to adopt both native (left-handed) and topologically mirror-image (right-handed) conformations in several coarse-grained and all-atom simulations.7,8,10–13 The molecular simulation with coarse-grained united-residue (UNRES) force field by Maisuradze et al. identified the mirror-image conformation of BdpA with very low population at room temperature.7 More recent replica exchange molecular dynamics (REMD) and structure-based simulations by Noel et al. further quantified the mirror-image conformation with the occupation of 5% among folded structures of BdpA.8 According to its small population, the mirror-image conformation of BdpA has been regarded as a misfolded state by the authors. In addition, both previous replica exchange Monte Carlo (REMC) simulation by Yang et al.14 and more recent molecular dynamics (MD) simulation enhanced by dihedral-biased tempering method9 indicated that another α-helix bundle protein, α3W (67 residues) which has natively left-handed topology, could also misfold to non-native mirror-image conformation with observable population. The common observation of the mirror-image conformation in the simulations of multiple protein systems and implemented with various force fields suggests that it should be not the force field artifact.

Understanding the molecular mechanism underlying protein folding and (possible) misfolding is of fundamental importance to biology. The difficulty of experimental technologies in distinguishing the fold and misfolded states of abovementioned proteins and their respective transient states in the folding transition pathways, however, limits our approach to explore the folding/misfolding mechanism of these proteins. All-atom MD simulation can essentially provide detailed dynamic information of protein motion, but the low efficiency in sampling high-dimensional and rugged configuration space of protein makes conventional MD simulation very computationally expensive.15 Enhanced sampling methods are thus often used to improve the sampling efficiency of MD simulation, ensuring that the simulation can sufficiently sample protein configuration space.9,16–22 The associated free energy landscape analysis of simulation data can identify and energetically characterize protein folding/misfolding pathway.

Recently, we used integrated-tempering-sampling molecular dynamics (hereinafter referred to as ITS-MD) simulation23,24 to systematically investigate the folding and misfolding of β-hairpin systems.3,5,25–27 The estimated folding kinetics and thermodynamics parameters (e.g., the folding time and individual backbone hydrogen bond stability of TRPZIP2, the cold denaturation tendency of five β-hairpins of MrH1, MrH4a, MrH3a, MrH3b, and MrH4b) are in well agreement with experimental data, indicating the validity of the simulation methodology in describing β-hairpin folding.3,28–30 More importantly, the misfolded states observed in most of β-hairpin systems are consistent with those reported by aforementioned simulation studies from other research groups.1,2,6 The detailed free energy landscape analysis sheds light on the mechanism of β-hairpin misfolding which is induced by the alterable turn configuration and its competition with the cross-strand hydrophobic side chain interactions.25 In the presence of strong β-turn-promoting sequence, the turn formation precedes over the formation of other two structural elements (the packing of cross-strand hydrophobic core cluster and backbone hydrogen bond assembling). On the other hand, the strong β-turn-promoting sequence might also have high risk for the turn adopting alternative conformations of which the native configuration does not but the non-native one does conflict with the packing of hydrophobic core cluster. As a result, the native turn configuration collaborates with the hydrophobic core to form correctly folded structure whereas the non-native turn configuration interrupts the packing of the hydrophobic core to reach misfolded structure, especially for protein system containing weak hydrophobic interactions. In the presence of weak β-turn-promoting sequence and strong hydrophobic residues on the strand, the turn formation becomes the rate-limiting step and the folding process is initiated by the packing of hydrophobic core cluster, which strongly limits the approach to β-hairpin misfolding. For α-helix bundle proteins, although previous studies suggested that right-handed conformation is dominated by charge interactions,9,31 how charge interactions take into effect in the folding process to help the protein find a routine to target structure and what is the interplay between the charge interactions and the driving force for protein structure collapse are still unclear.

In this article, we utilized the same enhanced sampling simulation method (ITS-MD) to simulate the folding of a 67-residue α-helix bundle protein, α3W (PDB code: 1LQ7 (ref. 32)). In comparison to other α-helix bundle analogues with relatively smaller sizes such as BdpA, the simulation studies on α3W are quite few, some of which have been performed using coarse-grained models.9,33,34 The detailed folding mechanism of the protein is thus not clear yet. The present all-atom simulation captures not only low-energy native-like (left-handed) and misfolded mirror-image (right-handed) conformations but also various high-energy configurations along the folding transition pathway back and forth in individual trajectories. The captured mirror-image conformation has the same topology as that reported in previous simulations of α3W.9,14 With sufficient protein folding/misfolding events collected in simulation, the folding and misfolding pathways of α3W can be identified through the free energy landscape analysis. It is observed that the folding pathway of α3W is mainly dominated by inter-helical hydrophobic interactions whereas the misfolding is also highly affected by inter-helical charge interactions. In addition, the present study suggests that the misfolding of mirror-image conformation should be easier to occur for α-helix bundle protein containing less hydrophobic but more charged residues. Such effects of intra-protein hydrophobic interaction strength on the misfolding tendency can be also seen for β-hairpin system.25 Therefore, the present study together with our previous studies on β-hairpins3,5,25 provide comprehensive, atomic-level picture of the folding and misfolding of α-helix bundle and β-hairpin.

2. Computational methods

All MD simulations were performed using AMBER11 package.35 AMBER FF03 all-atom force field36 was used to model protein and the recently improved generalized Born (GB) implicit solvent model, GB-Neck2 (igb = 8),37 was used to model solvent environment. The salt concentration was set to 0 and the default surface tension was 0.005 kcal mol−1 Å−2. The SHAKE algorithm38 with a relative geometric tolerance of 10−5 was used to constrain all chemical bonds. No non-bonded cutoff was used in any simulations.

Multiple independent simulation trajectories were performed. In each trajectory, the initial extended structure of the protein without residual secondary and tertiary structure elements was first subjected to 2500 steps of minimization. Then the system temperature was increased upon 300 K using velocity rearrangement from a Maxwell–Boltzmann distribution. After that, the system temperature was maintained at 300 K using the weak-coupling algorithm with a coupling time constant of 0.5 ps.39 The long-time equilibrium simulation (production run) was performed with a step size of 0.002 ps and at the constant temperature of 300 K, but with a potential modified via the ITS method to encourage thorough sampling over a large potential energy range. The calculation data were collected every 2.0 ps and all data from the production run were used for data analysis.

The details of ITS method have been described previously.23,24 Briefly, a modified potential energy is obtained from an integration function over temperature:

 
image file: c4ra14265a-t1.tif(1)
where image file: c4ra14265a-t2.tif (kB is the Boltzmann constant and T0 is the temperature of system), V is the original potential function of the system under study. βk is a series of temperatures that cover both low and high temperatures around the target one. The form of f(β) is taken as a sum of delta functions:
 
image file: c4ra14265a-t3.tif(2)

Then the integration in eqn (1) could be written as a summation over a series of discrete temperature values image file: c4ra14265a-t4.tif. Compared to conventional MD simulation, the sampled energy range in ITS-MD simulation could be greatly expanded depending upon the range of βk used. In the present study, 300 temperatures (βk), evenly distributed in the range of 270–390 K, were used. A preliminary iteration process was employed to obtain converged values of discrete nk's, which were then used in the long-time production run to evenly sample the entire energy range. After the data were collected in production run using the modified potential with the set of f(βk)'s, the thermodynamic properties of the simulated system at a desired temperature β were then calculated by reweighting of each term by timing a factor image file: c4ra14265a-t5.tif.

3. Results

3.1 One-dimensional free energy profile indicating the presence of both native and mirror-image conformations of α3W

Twelve independent trajectories starting from diverse extended structures of α3W were run in order to obtain converged simulation results. Each trajectory lasted around 600 ns and the accumulated simulation time was around 7 μs. Fig. 1 shows time series of protein backbone heavy-atom root-mean-square deviation (RMSD) with respect to the NMR native structure of α3W in a typical trajectory. One can see that reversible folding (RMSD ≤ 4.0 Å) and unfolding events can be achieved in individual simulation trajectory. The sampled structures most fitting the NMR structure can reach the backbone heavy-atom RMSD of 1.78 Å (see left panel of Fig. 1, the helices of H1–H3 are ordered from N- to C-terminal).
image file: c4ra14265a-f1.tif
Fig. 1 (Left panel) Comparison of the simulated structure with smallest backbone heavy-atom RMSD values (blue) and the NMR structure (red) of α3W. (Right panel) Time series of the backbone heavy-atom RMSD value in a typical trajectory of α3W.

The number of folding events is accumulated to ∼28 in all trajectories. The sufficient sampling of protein folding events allows us to statistically analyze the weight of various structures including the unfolded, transient, and folded structures along protein folding pathway. The one-dimensional free energy profile as a function of backbone heavy-atom RMSD was calculated at room temperature (Fig. 2(a)). As shown in this figure, the local free energy minimum at small RMSD position (∼3.90 Å) has the lowest free energy, indicating that the native-like (left-handed) state of α3W is the most populated. In addition, another local minimum with higher free energy is also presented at RMSD ≈ 9.2 Å, corresponding to the mirror-image (right-handed) state. The free energy difference between the native and misfolded states is about −0.37 kcal mol−1, making the former state roughly 1.86 times more populated than the latter state. The free energy profile depicted as a function of another reaction coordinate, the fraction of native contacts of α3W, also indicates that the native state is more thermodynamically favorable than the misfolded state (the free energy difference is −0.88 kcal mol−1, Fig. 2(b)). In addition, a free energy barrier needs to be overcome for the transition from the misfolded state to native state (1.40 kcal mol−1 in Fig. 2(a) and 0.628 kcal mol−1 in Fig. 2(b)).


image file: c4ra14265a-f2.tif
Fig. 2 One-dimensional free energy profiles as the function of (a) backbone heavy-atom RMSD value with respect to the NMR structure and (b) the fraction of native contacts of α3W at room temperature. The representative structures of native left-handed (RMSD ≤ 4.0 Å) and misfolded right-handed (RMSD ≈ 9.2 Å) conformations are shown at left and right sides, and their locations in the free energy profiles are pointed out by dash arrows, respectively.

To illuminate clearly the structure difference of the two states, we next calculated their contact maps using the contact map analysis (CMA) server in SPACE tools.40 One can see from Fig. 3 that the main difference is in the residue–residue contacts between the helix H3 and the other two helices (see the regions circled in Fig. 3). The observed mirror-image structure of α3W here has similar geometric features as that in previous explicit solvent MD simulation of the same protein:9 both mirror-image structures have similar chirality (opposite to the native structure) and have less side chain-side chain contacts among hydrophobic residues than the native structure.


image file: c4ra14265a-f3.tif
Fig. 3 The contact maps of (a) left-handed and (b) right-handed structures of α3W. The contact map is calculated by the contact map analysis (CMA) server in SPACE tools.40

It is worth noting that the use of ITS-MD simulation can save the computational resource in comparison to other enhanced sampling methods. In our previous study on an Ala–Pro peptide,41 we found that ITS-MD has an apparently higher efficiency than other bias potential and generalized ensemble methods, e.g., accelerated molecular dynamics (AMD) and replica-exchange molecular dynamics (REMD) simulations, in generating the thermodynamics parameters for the trans and cis interstate transition of the peptide. The present simulation ran 12 independent trajectories and each trajectory lasted ∼600 ns. To achieve converged free energy profiles for a smaller α-helix bundle protein, engrailed homeodomain protein (EnHD, 54 residues), Joshi and co-workers performed implicit solvent REMD simulations of which either 30 parallel trajectories (replicas) covering the temperature range of 286–373 K (ref. 42) or 26 replicas covering the temperature range of 292–367 K (ref. 43) were implemented and the simulation per replica was lasted for 600 ns–2.0 μs. The required computational resource and simulation time are greater than those in the present study. In addition, the coarse-grained canonical and multiplexed REMD simulation on the 46-residue BdpA protein used 80 replicas to cover the temperature range of 250–500 K.7 The use of explicit solvent REMD to study the folding of protein is even more computationally expensive, e.g., a total of 64 replicas with the temperature range of 287–643 K were used by Noel et al. and 82 replicas with the temperature range of 277–548 K were used by Garcia et al. in their respective folding simulations on BdpA protein.8,44

3.2 Two-dimensional free energy landscapes indicating the folding/misfolding mechanisms of native and mirror-image conformations of α3W

The folding of protein involves the formation of both secondary and tertiary structures. Since the native left-handed and misfolded right-handed conformations of α3W share same secondary structure elements (α-helices of H1–H3 from N- to C-terminal), the molecular interactions which distinguish the two conformations are tertiary interactions (side chain-side chain interactions). The tertiary interactions can be divided into hydrophobic interactions from non-polar residues, favorite charge interactions from oppositely charged residues (salt bridges), and unfavorite charge interactions from residues carrying same types of charges. All of these interactions within individual subdomains (two subdomains formed by connecting helices, e.g., H1–H2, H2–H3, and one subdomain formed between two remote helices, H1–H3) are listed in Table 1 for both conformations. In comparison to the left-handed conformation, the right-handed conformation of α3W contains similar amount of favorite salt bridges but less unfavorite charge interactions, indicating the intrinsic propensity of right-handed conformation.
Table 1 The inter-helical hydrophobic contacts, favorite and unfavorite charge interactions formed in individual subdomains in native left-handed and misfolded right-handed conformations of α3W
Molecular interactions Native (left-handed) conformation Misfolded (right-handed) conformation
H1–H2 H1–H3 H2–H3 H1–H2 H1–H3 H2–H3
Inter-helical hydrophobic interactions V4–I41 V4–V50 I27–I64 V4–L44 L7–V53 I27–I64
V4–L44 L7–V50 I27–L67 A6–L44 L7–V57 I27–L67
L7–L37 L7–V53 L30–L60 L7–L37 V11–V53 W34–V57
L7–I41 L7–V57 L30–I64 L7–I41 V11–V57 W34–L60
L7–L44 V11–V57 L30–L67 L7–L44 V11–L60 W34–I64
V11–W34 L14–V57 W34–V57 L14–L30 L14–L60 L37–V57
V11–L37 L14–L60 W34–L60 L14–W34 L14–I64 L37–L60
V11–I41 L14–I64 L37–V53 L14–L37 V18–L60 I41–V53
L14W34 V18–I64 L37–V57 V18–I27 V18–I64 I41–V57
V18–I27 L21–I64 L37–L60 V18–L30 V18–L67
V18–L30 L21–L67 I41–V53 L21–I27
V18–W34 L44–V53
Inter-helical favorite charge interactions (salt bridges) E15–K38 R3–E54 E29–K59 K17–E29 R3–E54 K38–E61
K10–E61 K33–E56 K12–E56
K17–E61 K40–E56 E15–K59
K19–E63
Inter-helical unfavorite charge interactions K12–K38 R3–K51 E29–E63 K10–K33 E8–E56
K19–K31 K10–K58 K33–K59 K17–R26 E15–E63
K40–K52


To understand the interplay of the formation of secondary and tertiary structures in the folding/misfolding pathways of α3W, we used the total number of backbone hydrogen bonds formed (NHB) and the number of inter-helical side chain contacts (CNat) to depict two-dimensional free energy landscapes for both native and mirror-image conformations. The former reaction coordinate is often used to describe secondary structure and the latter corresponds to tertiary structure. It is noteworthy that the detailed backbone hydrogen bonds are similar but the components of inter-helical side chain contacts are different for the two conformations.

As shown in Fig. 4, three distinct local minima exist in these profiles. A large fraction of secondary structure contents are formed even in local minima containing fewest side chain contacts. Thus, the three local minima could be treated as two partially folded/misfolded states (PL1 and PL2 in Fig. 4(a) corresponding to the partially folded left-handed states and PR1 and PR2 in Fig. 4(b) corresponding to the partially misfolded right-handed states) and one folded (F) or misfolded (M) state. The connection of the three distinct states in each profile constructs the lowest free energy folding or misfolding pathway for protein. One can see that the number of backbone hydrogen bonds increases and reaches the maximum in very early stage of the folding of α3W (PL1 or PR1 state) where only a small portion of inter-helical side chain contacts are formed. The further folding of α3W involves the formation of more side chain contacts while the secondary structure is fully formed. Therefore, for either native or misfolded conformation, the formation of secondary structures precedes over that of tertiary structure, consistent with the framework mechanism: in the folding process of protein, its secondary structures are formed incipiently and independently of tertiary structure, then the formed secondary structure elements coalesce into native structure.45 This observation is in well agreement with previous large-scale Monte Carlo simulation study of α3W.33


image file: c4ra14265a-f4.tif
Fig. 4 Two-dimensional free energy landscapes as the function of CNat and NHB for (a) native left-handed and (b) misfolded right-handed conformations of α3W at room temperature. The contours are spaced at intervals of 0.5kBT.

The fractions of hydrophobic and charge interactions in each distinct state along the folding and misfolding pathways of α3W, respectively, were calculated. One can see from Table 2 that the fraction of hydrophobic interactions increases gradually in the transition pathways of both native and misfolded conformations. For left-handed conformation, the fraction of charge interactions is always smaller than that of hydrophobic interactions in all three states, suggesting that hydrophobic interactions but not charge interactions largely contribute for the folding of left-handed conformation. In contrast, although only a small fraction of charge interactions are formed in the very early stage of the formation of right-handed conformation (PR1 state), most of charge interactions are formed in the middle and all of them are formed at the end of the misfolding pathway, strongly suggesting that charge interactions also play an important role in the misfolding of α3W to right-handed conformation. This observation is consistent with the speculation that charge interactions dominate the formation of right-handed conformation of α-helix bundle.9,31

Table 2 The average fraction of inter-helical hydrophobic and charge interactions formed in the three distinct states of the native and misfolded conformations of α3W, respectively
  Left-handed conformation Right-handed conformation
PL1 state PL2 state F state PR1 state PR2 state M state
Fraction of hydrophobic interactions 31.26% 43.50% 70.08% 32.07% 52.88% 79.16%
Fraction of charge interactions 18.50% 26.98% 35.66% 16.10% 65.10% 87.08%


Next, to distinguish the folding pathway of left-handed conformation from the misfolding pathway of right-handed conformation, we calculated the free energy landscape as the function of the fraction of all inter-helical side chain contacts for both conformations (QLeft and QRight). The local minima corresponding to the partially folded/misfolded and fully folded/misfolded states are also presented in Fig. 5. It is clear to be seen that the lowest free energy folding and misfolding pathways are roughly mirror symmetric along the diagonal axis. Therefore, the divergence of the folding of left-handed and the misfolding of right-handed conformations of α3W occurs in the very early stage of α3W folding process. The average number of inter-helical hydrophobic contacts (the main content of tertiary interactions) formed within each subdomain in the distinct states along the folding/misfolding pathways of left-handed and right-handed conformations was calculated and organized in Table 3. For both conformations, more hydrophobic contacts are formed in H1–H2 subdomain than in the other two subdomains in the early stage of α3W folding (PL1 or PR1 state) and the further folding leads to the formation of more hydrophobic contacts in all subdomains.


image file: c4ra14265a-f5.tif
Fig. 5 Two-dimensional free energy landscapes as the function of the fraction of inter-helical side chain contacts for left- and right-handed conformations (QLeft and QRight) at room temperature. The contours are spaced at intervals of 0.2kBT.
Table 3 The average number of inter-helical hydrophobic contacts formed in each subdomain for the three distinct states of the native and misfolded conformations of α3W, respectively
  Left native conformation Right non-native conformation
PL1 state PL2 state FL state PR1 state PR2 state FR state
NHC, H1–H2 5.25 5.53 8.25 4.98 6.42 10.35
NHC, H1–H3 2.84 6.50 8.52 2.88 3.02 5.84
NHC, H2–H3 2.85 3.17 7.75 3.68 5.76 6.28


3.3 Temperature dependence of the stability of native and mirror-image conformations

The molecular simulation with coarse-grained united-residue (UNRES) force field by Maisuradze et al. showed that the native and mirror-image conformations of BdpA protein display opposite temperature dependence in their stabilities: the population of the former conformation decreases whereas the population of the latter increases as the temperature is increased.7 The REMD and structure-based simulations by Noel et al. demonstrated that the populations of both conformations of BdpA are decreased with temperature in general and the decrease in population of native conformation is much severer than that of mirror-image conformation.8 Despite the discrepancy in the description of temperature dependence of mirror-image conformation, both simulation studies suggested that the mirror-image conformation of BdpA has strong temperature resistance than native conformation. To evaluate the temperature resistance of the conformations with native and mirror-image topologies of α3W, we calculated the average fraction of inter-helical side chain contacts formed for both conformations at different temperatures (the number is averaged on the basis of all structures sampled in the simulation). One can see from Fig. 6 that the contact fraction decreases with the temperature and the decreasing tendency of native conformation is more apparent than that of mirror-image conformation, suggesting that the former conformation has lower temperature resistance than the latter. As a result, the population difference between native and mirror-image conformations should be shortened as the temperature is increased.
image file: c4ra14265a-f6.tif
Fig. 6 The temperature dependence of the average fraction of inter-helical side chain interactions for left- and right-handed conformations of α3W.

To understand the different temperature resistance of the native and mirror-image conformations of α3W, we calculated the average fraction of inter-helical hydrophobic and charge interactions formed, respectively (Fig. 7). Both kinds of interactions have their fraction decreased with temperature in native conformation, corresponding to the low temperature resistance of this conformation. However, while the fraction of hydrophobic interactions keeps decreasing, the fraction of charge interactions in the non-native right-handed conformation keeps small but steady in a large temperature range. The small value of the fraction of charge interactions corresponds to the low population of non-native right-handed conformation and its steadiness explains the relatively strong temperature resistance of this conformation.


image file: c4ra14265a-f7.tif
Fig. 7 The temperature dependence of the average fraction of inter-helical hydrophobic and charge interactions for (a) left- and (b) right-handed conformations of α3W.

4. Discussion and conclusions

Recent molecular dynamics simulations suggested that protein misfolding could be a universal phenomenon among simple β-structured and α-helical proteins/polypeptides.1–9 The less population of misfolded structure in comparison to that of native structure, however, makes it difficult to be observed for experimental technologies, which leads to the difficulty in understanding the misfolding mechanism. Recently, we systematically investigated the folding and misfolding of β-hairpin using all-atom molecular dynamics simulation enhanced by integrated-tempering-sampling method, which gives reasonable explanation of β-hairpin misfolding.5,25–28 In the present study, to understand the folding and misfolding mechanism of α-helix bundle protein, we applied the same methodology on the folding simulation of a typical three-α-helix bundle, α3W.

In principle, the accuracy of theoretical simulation largely depends on the molecular force field used. The popular combination of AMBER FF03 force field and GB implicit solvent model, which has been widely used in many previous simulations on the members of α-helix bundle proteins,42,46–50 was used here to model protein amino acids and water solvent, respectively. Although it is well-known that FF03 favors α-helical conformations,51–54 these abovementioned simulations can still provide reasonable description on the folding of α-helix bundle proteins, with the calculated thermodynamics parameters (e.g., melting temperature and folding free energy) consistent with experimental data. Very recently, Simmerling and co-workers indicated that the molecular simulation combined with the improved generalized Born (GB) implicit solvent model, GB-Neck2 (igb = 8),37 can fold proteins with diverse topologies.55 The combination of AMBER FF03 force field and GB-Neck2 model in the present simulation can not only fold α3W into its NMR experimentally identified native (left-handed) structure but also capture a misfolded mirror-image (right-handed) structure “hidden” behind the native one which has the similar topology as that observed in previous explicit solvent MD simulation,9 indicating its efficiency in the sampling of protein configuration space. The former conformation occupies the largest fraction of the equilibrium conformational ensembles sampled by the simulation whereas the population of the latter is much smaller.

The detailed free energy landscape analysis indicates that the folding of α3W starts with the formation of all α-helical secondary structures and is then followed by the assembling (packing) of folded α-helices to construct the tertiary topology. The α-helix assembling which involves the formation of inter-helical side chain contacts e.g., by hydrophobic and charge interactions, however, could divaricate at the early stage of folding process, depending on which kinds of interaction makes dominant contribution in the structure assembling. The “correct” assembling which leads to native topology is solely dominated by inter-helical hydrophobic interactions whereas the “incorrect” assembling which leads to misfolded mirror-image structure is highly affected by not only inter-helical hydrophobic interactions but also charge interactions. The important role of hydrophobic interactions in folding native structure can be seen from the larger amount of hydrophobic interactions in native structure than those in mirror-image structure (Table 1) as well as the large fraction of hydrophobic interactions formed in the folding pathway (Table 2). Similarly, the large fraction of charge interactions in the misfolding pathway (Table 2) reflects its importance in α3W misfolding.

Considering the importance of hydrophobic interactions to folding and the importance of charge interactions to misfolding of α3W, one may speculate that the misfolding probability could be increased in the presence of less hydrophobic but more charged residues in amino acid sequence for α-helix bundle protein. Using α3W, BdpA, and another α-helix bundle protein α3D as examples, one can see that the hydrophobic and charged residues occupy 31.34% and 53.74% in the amino acid sequence of α3W, and the counterparts in BdpA occupy 41.30% and 26.08% and in α3D occupy 51.8% and 33.2%, respectively. Accordingly, the mirror-image structure of α3W has larger population than the mirror-image structure of BdpA (roughly 34.96% of folded structures of α3W are in mirror-image conformation as shown in the present study but the number is only 5% for BdpA in ref. 8). In addition, both the explicit solvent MD simulation by Zhang and Ma9 and our recent implicit solvent MD simulation50 showed that no mirror-image structure is adopted for α3D which has the largest fraction of hydrophobic residues in the sequence among the three proteins under study. Therefore, the misfolding probability of an α-helix bundle protein could be evaluated by its amino acid sequence information. It is worth noting that our previous simulation suggested that the folding mechanism of an α-helix bundle protein is also dependent on its sequence: α-helical proteins containing higher percentage of hydrophobic residues than charged ones fold via nucleation–condensation mechanism (concurrent formation of secondary and tertiary structures) whereas proteins having opposite tendency in amino acid composition more likely fold via framework mechanism (incipient and independent formation of secondary structures before tertiary interactions).50 In this scenario, both BdpA and α3D fold via nucleation–condensation mechanism whereas the folding of α3W follows framework mechanism.50 Therefore, one may further speculate that the cooperative folding (nucleation–condensation) probably decrease the probability of misfolding. Combining the present study and our previous studies on β-hairpins,5,23–26 we can find one common feature in the misfolding of α-helix bundle and β-hairpin: proteins containing stronger intra-protein hydrophobic interactions have smaller probability to misfold.

Acknowledgements

We thank the National Natural Science Foundation of China (Grant no. 21373258) and National Basic Research Program (Grant no. 2014CB910400) for financial support. The simulations were run at TianHe 1 supercomputer in Tianjin, and supercomputing facilities at The Texas Advanced Computing Center (TACC) at The University of Texas at Austin (Project ID: TG-MCB110130) and BlueBioU at Rice University.

References

  1. R. B. Best and J. Mittal, Proc. Natl. Acad. Sci. U. S. A., 2011, 108, 11087–11092 CrossRef CAS PubMed.
  2. A. Bhattacharya, R. B. Best and J. Mittal, Biophys. J., 2012, 103, 596–600 CrossRef CAS PubMed.
  3. Q. Shao, J. Shi and W. Zhu, J. Chem. Phys., 2013, 138, 085102 CrossRef PubMed.
  4. Q. Shao, L. Yang and Y. Q. Gao, J. Chem. Phys., 2009, 130, 195104 CrossRef PubMed.
  5. Q. Shao, L. Yang and Y. Q. Gao, J. Chem. Phys., 2011, 135, 235104 CrossRef PubMed.
  6. P. Kuhrova, A. De Simone, M. Otyepka and R. B. Best, Biophys. J., 2012, 102, 1897–1906 CrossRef CAS PubMed.
  7. G. G. Maisuradze, A. Liwo, S. Oldziej and H. A. Scheraga, J. Am. Chem. Soc., 2010, 132, 9444–9452 CrossRef CAS PubMed.
  8. J. K. Noel, A. Schug, A. Verma, W. Wenzel, A. E. Garcia and J. N. Onuchic, J. Phys. Chem. B, 2012, 116, 6880–6888 CrossRef CAS PubMed.
  9. C. Zhang and J. P. Ma, Proc. Natl. Acad. Sci. U. S. A., 2012, 109, 8139–8144 CrossRef CAS PubMed.
  10. J. Lee, A. Liwo and H. A. Scheraga, Proc. Natl. Acad. Sci. U. S. A., 1999, 96, 2025–2030 CrossRef CAS.
  11. A. Irback, F. Sjunnesson and S. Wallin, Proc. Natl. Acad. Sci. U. S. A., 2000, 97, 13614–13618 CrossRef CAS PubMed.
  12. K. Kachlishvili, G. G. Maisuradze, O. A. Martin, A. Liwo, J. A. Vila and H. A. Scheraga, Proc. Natl. Acad. Sci. U. S. A., 2014, 111, 8458–8463 CrossRef CAS PubMed.
  13. K. A. Olszewski, A. Kolinski and J. Skolnick, Proteins Struct. Funct. Genet., 1996, 25, 286–299 CAS.
  14. J. S. Yang, W. W. Chen, J. Skolnick and E. I. Shakhnovich, Structure, 2007, 15, 53–63 CrossRef CAS PubMed.
  15. J. L. Klepeis, K. Lindorff-Larsen, R. O. Dror and D. E. Shaw, Curr. Opin. Struct. Biol., 2009, 19, 120–127 CrossRef CAS PubMed.
  16. S. Kumar, D. Bouzida, R. H. Swendsen, P. A. Kollman and J. M. Rosenberg, J. Comput. Chem., 1992, 13, 1011–1021 CrossRef CAS.
  17. S. Kumar, J. M. Rosenberg, D. Bouzida, R. H. Swendsen and P. A. Kollman, J. Comput. Chem., 1995, 16, 1339–1350 CrossRef CAS.
  18. U. H. E. Hansmann, Chem. Phys. Lett., 1997, 281, 140–150 CrossRef CAS.
  19. G. Bussi, A. Laio and M. Parrinello, Phys. Rev. Lett., 2006, 96, 090601 CrossRef.
  20. A. Mitsutake and Y. Okamoto, Chem. Phys. Lett., 2000, 332, 131–138 CrossRef CAS.
  21. F. G. Wang and D. P. Landau, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2001, 64, 056101 CrossRef CAS.
  22. F. G. Wang and D. P. Landau, Phys. Rev. Lett., 2001, 86, 2050–2053 CrossRef CAS.
  23. Y. Q. Gao, J. Chem. Phys., 2008, 128, 064105 CrossRef PubMed.
  24. Q. Shao, W. L. Zhu and Y. Q. Gao, J. Phys. Chem. B, 2012, 116, 13848–13856 CrossRef CAS PubMed.
  25. Q. Shao, J. Wang, J. Shi and W. Zhu, J. Chem. Phys., 2013, 139, 165103 CrossRef PubMed.
  26. Q. Shao and Y. Q. Gao, J. Chem. Theory Comput., 2010, 6, 3750–3760 CrossRef CAS.
  27. Q. Shao, H. Wei and Y. Q. Gao, J. Mol. Biol., 2010, 402, 595–609 CrossRef CAS PubMed.
  28. L. J. Yang, Q. Shao and Y. Q. Gao, J. Phys. Chem. B, 2009, 113, 803–808 CrossRef CAS PubMed.
  29. A. J. Maynard, G. J. Sharman and M. S. Searle, J. Am. Chem. Soc., 1998, 120, 1996–2007 CrossRef CAS.
  30. R. B. Dyer, S. J. Maness, S. Franzen, R. M. Fesinmeyer, K. A. Olsen and N. H. Andersen, Biochemistry, 2005, 44, 10406–10415 CrossRef CAS PubMed.
  31. J. W. Bryson, J. R. Desjarlais, T. M. Handel and W. F. DeGrado, Protein Sci., 1998, 7, 1404–1414 CrossRef CAS PubMed.
  32. Q. H. Dai, C. Tommos, E. J. Fuentes, M. R. A. Blomberg, P. L. Dutton and A. J. Wand, J. Am. Chem. Soc., 2002, 124, 10952–10953 CrossRef CAS PubMed.
  33. J. H. Meinke and U. H. E. Hansmann, J. Comput. Chem., 2009, 30, 1642–1648 CrossRef CAS PubMed.
  34. C. Czaplewski, S. Kalinowski, A. Liwo and H. A. Scheraga, J. Chem. Theory Comput., 2009, 5, 627–640 CrossRef CAS PubMed.
  35. D. A. Case, T. A. Darden, T. E. Cheatham III, C. L. Simmerling, J. Wang, R. E. Duke, R. Luo, R. C. Walker, W. Zhang, K. M. Merz, B. Roberts, B. Wang, S. Hayik, A. Roitberg, G. Seabra, I. Kolossváry, K. F. Wong, F. Paesani, J. Vanicek, J. Liu, X. Wu, S. R. Brozell, T. Steinbrecher, H. Gohlke, Q. Cai, X. Ye, J. Wang, M.-J. Hsieh, G. Cui, D. R. Roe, D. H. Mathews, M. G. Seetin, C. Sagui, V. Babin, T. Luchko, S. Gusarov, A. Kovalenko and P. A. Kollman, AMBER11, University of California, San Francisco, 2010 Search PubMed.
  36. Y. Duan, C. Wu, S. Chowdhury, M. C. Lee, G. M. Xiong, W. Zhang, R. Yang, P. Cieplak, R. Luo, T. Lee, J. Caldwell, J. M. Wang and P. Kollman, J. Comput. Chem., 2003, 24, 1999–2012 CrossRef CAS PubMed.
  37. H. Nguyen, D. R. Roe and C. Simmerling, J. Chem. Theory Comput., 2013, 9, 2020–2034 CrossRef CAS.
  38. J. P. Ryckaert, G. Ciccotti and H. J. C. Berendsen, J. Comput. Phys., 1977, 23, 327–341 CrossRef CAS.
  39. H. J. C. Berendsen, J. P. M. Postma, W. F. Vangunsteren, A. Dinola and J. R. Haak, J. Chem. Phys., 1984, 81, 3684–3690 CrossRef CAS PubMed.
  40. V. Sobolev, E. Eyal, S. Gerzon, V. Potapov, M. Babor, J. Prilusky and M. Edelman, Nucleic Acids Res., 2005, 33, W39–W43 CrossRef CAS PubMed.
  41. L. J. Yang, Q. Shao and Y. Q. Gao, J. Chem. Phys., 2009, 130, 124111 CrossRef PubMed.
  42. S. Koulgi, U. Sonavane and R. Joshi, J. Mol. Graphics Modell., 2010, 29, 481–491 CrossRef CAS PubMed.
  43. V. Jani, U. B. Sonavane and R. Joshi, J. Mol. Model., 2014, 20, 2283 CrossRef PubMed.
  44. A. E. Garcia and J. N. Onuchic, Proc. Natl. Acad. Sci. U. S. A., 2003, 100, 13898–13903 CrossRef CAS PubMed.
  45. P. S. Kim and R. L. Baldwin, Annu. Rev. Biochem., 1990, 59, 631–660 CrossRef CAS PubMed.
  46. H. X. Lei, C. Wu, Z. X. Wang, Y. Q. Zhou and Y. Duan, J. Chem. Phys., 2008, 128, 235105 CrossRef PubMed.
  47. H. X. Lei and Y. Duan, J. Phys. Chem. B, 2007, 111, 5458–5463 CrossRef CAS PubMed.
  48. Q. Shao and Y. Q. Gao, J. Chem. Phys., 2011, 135, 135102 CrossRef PubMed.
  49. H. X. Lei, C. Wu, H. G. Liu and Y. Duan, Proc. Natl. Acad. Sci. U. S. A., 2007, 104, 4925–4930 CrossRef CAS PubMed.
  50. Q. Shao, J. Phys. Chem. B, 2014, 118, 5891–5900 CrossRef CAS PubMed.
  51. S. Ono, N. Nakajima, J. Higo and H. Nakamura, J. Comput. Chem., 2000, 21, 748–762 CrossRef CAS.
  52. M. H. Zaman, M. Y. Shen, R. S. Berry, K. F. Freed and T. R. Sosnick, J. Mol. Biol., 2003, 331, 693–711 CrossRef CAS.
  53. J. Higo, N. Ito, M. Kuroda, S. Ono, N. Nakajima and H. Nakamura, Protein Sci., 2001, 10, 1160–1171 CrossRef CAS PubMed.
  54. N. Kamiya, J. Higo and H. Nakamura, Protein Sci., 2002, 11, 2297–2307 CrossRef CAS PubMed.
  55. H. Nguyen, J. Maier, H. Huang, V. Perrone and C. Simmerling, J. Am. Chem. Soc., 2014, 136, 13959–13962 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.