arxiv:cond-mat/ v1 [cond-mat.soft] 19 Mar 2001

Similar documents
arxiv:cond-mat/ v1 [cond-mat.soft] 5 May 1998

Identifying the Protein Folding Nucleus Using Molecular Dynamics

Local Interactions Dominate Folding in a Simple Protein Model

Master equation approach to finding the rate-limiting steps in biopolymer folding

Folding of small proteins using a single continuous potential

arxiv:cond-mat/ v1 [cond-mat.soft] 23 Mar 2007

arxiv:cond-mat/ v1 [cond-mat.soft] 16 Nov 2002

The role of secondary structure in protein structure selection

FREQUENCY selected sequences Z. No.

The kinetics of protein folding is often remarkably simple. For

Many proteins spontaneously refold into native form in vitro with high fidelity and high speed.

arxiv: v1 [cond-mat.soft] 22 Oct 2007

Protein Folding Prof. Eugene Shakhnovich

Modeling protein folding: the beauty and power of simplicity Eugene I Shakhnovich

Outline. The ensemble folding kinetics of protein G from an all-atom Monte Carlo simulation. Unfolded Folded. What is protein folding?

A simple model for calculating the kinetics of protein folding from three-dimensional structures

Short Announcements. 1 st Quiz today: 15 minutes. Homework 3: Due next Wednesday.

arxiv:cond-mat/ v1 7 Jul 2000

A new combination of replica exchange Monte Carlo and histogram analysis for protein folding and thermodynamics

DesignabilityofProteinStructures:ALattice-ModelStudy usingthemiyazawa-jerniganmatrix

arxiv:cond-mat/ v1 2 Feb 94

Effective stochastic dynamics on a protein folding energy landscape

The starting point for folding proteins on a computer is to

Temperature dependence of reactions with multiple pathways

It is not yet possible to simulate the formation of proteins

To understand pathways of protein folding, experimentalists

Long Range Moves for High Density Polymer Simulations

Protein Folding In Vitro*

Pathways for protein folding: is a new view needed?

Universal correlation between energy gap and foldability for the random energy model and lattice proteins

Monte Carlo simulations of polyalanine using a reduced model and statistics-based interaction potentials

Simulation of mutation: Influence of a side group on global minimum structure and dynamics of a protein model

PROTEIN EVOLUTION AND PROTEIN FOLDING: NON-FUNCTIONAL CONSERVED RESIDUES AND THEIR PROBABLE ROLE

Guessing the upper bound free-energy difference between native-like structures. Jorge A. Vila

Nature of the transition state ensemble for protein folding

Effect of Sequences on the Shape of Protein Energy Landscapes Yue Li Department of Computer Science Florida State University Tallahassee, FL 32306

Molecular dynamics simulations of anti-aggregation effect of ibuprofen. Wenling E. Chang, Takako Takeda, E. Prabhu Raman, and Dmitri Klimov

arxiv:chem-ph/ v1 11 Nov 1994

Stretching lattice models of protein folding

Sparsely populated folding intermediates of the Fyn SH3 domain: Matching native-centric essential dynamics and experiment

Toward an outline of the topography of a realistic proteinfolding funnel

A Minimal Physically Realistic Protein-Like Lattice Model: Designing an Energy Landscape that Ensures All-Or-None Folding to a Unique Native State

PROTEIN FOLDING THEORY: From Lattice to All-Atom Models

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 1 Oct 1999

Protein Folding Pathways and Kinetics: Molecular Dynamics Simulations of -Strand Motifs

Visualizing folding of proteins (1 3) and RNA (2) in terms of

Lecture 34 Protein Unfolding Thermodynamics

Elucidation of the RNA-folding mechanism at the level of both

Paul Sigler et al, 1998.

= (-22) = +2kJ /mol

1 of 31. Nucleation and the transition state of the SH3 domain. Isaac A. Hubner, Katherine A. Edmonds, and Eugene I. Shakhnovich *

A molecular dynamics investigation of the kinetic bottlenecks of the hpin1 WW domain. II: simulations with the Go model

Exploring protein-folding ensembles: A variable-barrier model for the analysis of equilibrium unfolding experiments

Simulating Folding of Helical Proteins with Coarse Grained Models

Comparison of Two Optimization Methods to Derive Energy Parameters for Protein Folding: Perceptron and Z Score

Folding pathway of a lattice model for protein folding

Computer simulation of polypeptides in a confinement

Phase behavior of a lattice protein model

Influence of temperature and diffusive entropy on the capture radius of fly-casting binding

Prediction of protein-folding mechanisms from free-energy landscapes derived from native structures

Scattered Hammond plots reveal second level of site-specific information in protein folding: ( )

Engineering of stable and fast-folding sequences of model proteins

Random heteropolymer adsorption on disordered multifunctional surfaces: Effect of specific intersegment interactions

Evolution of Model Proteins on a Foldability Landscape

Pair potentials for protein folding: choice of reference states and sensitivity of predicted native states to variations in the interaction schemes

Protein Folding & Stability. Lecture 11: Margaret A. Daugherty. Fall How do we go from an unfolded polypeptide chain to a

Effects of Crowding and Confinement on the Structures of the Transition State Ensemble in Proteins

Lecture 2 and 3: Review of forces (ctd.) and elementary statistical mechanics. Contributions to protein stability

Nucleation and the Transition State of the SH3 Domain

Folding of a Small Helical Protein Using Hydrogen Bonds and Hydrophobicity Forces

Lecture 11: Protein Folding & Stability

Protein Folding & Stability. Lecture 11: Margaret A. Daugherty. Fall Protein Folding: What we know. Protein Folding

arxiv:q-bio/ v1 [q-bio.bm] 11 Jan 2004

Evolution of functionality in lattice proteins

Chapter 12 What Is Temperature? What Is Heat Capacity?

The protein folding problem consists of two parts:

Determination of Barrier Heights and Prefactors from Protein Folding Rate Data

Research Paper 577. Correspondence: Nikolay V Dokholyan Key words: Go model, molecular dynamics, protein folding

The folding mechanism of larger model proteins: Role of native structure

Protein structure forces, and folding

Computer simulations of protein folding with a small number of distance restraints

Energy Landscape and Global Optimization for a Frustrated Model Protein

Two-State Folding over a Weak Free-Energy Barrier

Thermodynamics. Entropy and its Applications. Lecture 11. NC State University

arxiv:physics/ v1 [physics.bio-ph] 7 Mar 2002

THE TANGO ALGORITHM: SECONDARY STRUCTURE PROPENSITIES, STATISTICAL MECHANICS APPROXIMATION

Cooperativity, Local-Nonlocal Coupling, and Nonnative Interactions in Protein Folding

arxiv:cond-mat/ v1 6 Nov 2002

Molecular Interactions F14NMI. Lecture 4: worked answers to practice questions

Designing refoldable model molecules

Simulating disorder order transitions in molecular recognition of unstructured proteins: Where folding meets binding

Self-consistently optimized energy functions for protein structure prediction by molecular dynamics (protein folding energy landscape)

The topomer search model: A simple, quantitative theory of two-state protein folding kinetics

Accounting for Protein-Solvent Contacts Facilitates Design of Nonaggregating Lattice Proteins

Entropic barriers, transition states, funnels, and exponential protein folding kinetics: A simple model

Single-molecule spectroscopy has recently emerged as a powerful. Effects of surface tethering on protein folding mechanisms

The chemical environment exerts a fundamental influence on the

Folding with Downhill Behavior and Low Cooperativity of Proteins

BCHS 6229 Protein Structure and Function. Lecture 3 (October 18, 2011) Protein Folding: Forces, Mechanisms & Characterization

Macromolecule Stability Curves

Transcription:

Modeling two-state cooperativity in protein folding Ke Fan, Jun Wang, and Wei Wang arxiv:cond-mat/0103385v1 [cond-mat.soft] 19 Mar 2001 National Laboratory of Solid State Microstructure and Department of Physics, Nanjing University, Nanjing 210093, China (February 19, 2001) Abstract A protein model with the pairwise interaction energies varying as local environment changes, i.e., including some kinds of collective effect between the contacts, is proposed. Lattice Monte Carlo simulations on the thermodynamical characteristics and free energy profile show a well-defined two-state behavior and cooperativity of folding for such a model. As a comparison, related simulations for the usual Gō model, where the interaction energies are independent of the local conformations, are also made. Our results indicate that the evolution of interactions during the folding process plays an important role in the two-state cooperativity in protein folding. PACS number: 87.15 Aa, 87.15 Cc, 87.15 He Typeset using REVTEX 1

Due to the developments of experimental methods and theoretical models, many achievements of protein folding have been made recently [1]. A protein can fold itself to its uniquely well-defined native structure in a biologically short time, regardless of the huge number of possible conformations, showing a highly cooperatively kinetic behavior. It is now clear that the cooperativity of folding may result from the backbone hydrogen bonding, sidechain packing and hydrophobic interactions, among them the hydrophobic interactions are believed to be the dominant driving force for folding [2]. For many small single-domain proteins or lattice proteinlike models, there is a two-state behavior between the unfolded states and the folded native one [3,4]. Recently, Chan and Kaya [5] indicated that according to the calorimetric criterion, which is widely used in experiments as a condition for two-state folding, popular lattice models, e.g., two-letter HP and 20-letter MJ models, are far from two-state models. Thismay beduetosome flawed assumptions inthepotentialfunctions usedinthese models. Lattice models usually use statistical potential functions extracted from the pairing frequencies of 20 kinds of amino acids in databases of protein structures [6]. Although these knowledge-based potentials may be a good approximation to the relative strength of interactions between the residues in the native state, they provide no information about how the interactions evolve during the folding. For computational convenience, a common assumption in lattice models is that the interactions are additive, and they are the same during the folding as in the native state. This means that the interaction energies are conformation-independent. Clearly this is not relevant to the experimental situation [7]. In fact, as Dill pointed out [8], the thermodynamic additivity principle which is widely used in chemistry may be unsuitable in biochemistry. Some recent experiments also indicated that the transition state is an expanded version of the native state, where the majority of interactions are partially formed [9], and their strengths are different from those in the native state (with Φ < 1). That is, these interactions depend on the conformations [7], especially the local structures around the contacts as emphasized recently in Ref. [10]. Previously, the non-additivity was built in a lattice model for packing effects [11]; and the hydrophobic force depending on the local density of peptide atoms was also taken into account in an off-lattice 2

model [12]. Studies on these models show that the introduction of the non-additivity is significiant, but the two-state cooperativity of these models is not checked, and the effects of the non-additivity on thermodynamics and kinetics of folding need to be further studied. In this paper, we develop a refined Gō model where the pairwise interaction energies vary as the local environment changes, i.e., some kinds of collective effects between contacts are introduced. Our purpose here is to study the two-state cooperativity of protein folding and its physical origin with such a model. Our results give a general picture about how the conformation-dependent interactions affect the folding kinetics, which is consistent with the phenomenological explanation based on experimental results. We model a polypeptide chain as a self-avoiding chain on a cubic lattice. A contact is formed if two residues are space adjacent but not sequence adjacent. If two residues form a contact as the same as in the native state, we call this contact a native contact, otherwise a non-native contact. Following the Gō model [13], only native contacts are considered to contribute to the total energy. Different from the Gō model, we assume that the interaction energies between residues are conformation-dependent, and vary with changes of the local environment. To achieve this, we introduce a parameter S to describe the degree for a residue being ordered relative to the native state. For the i-th residue in a certain conformation, its degree of order S i is defined as S i = z i /z nat i, (1) where z i is the number of native contacts in this conformation, z nat i is the number of contacts formed in the native state. Obviously, S i varies between 0 (the i-th residue being fully disordered) and 1 (being fully ordered). Thus, the interaction energy between residues i and j, B ij =-(S i +S j )ε/2 is defined, where S i and S j are the degrees of order for residues i and j, respectively. ε is the unit of energy and is set to be 1 in this work. The total energy of the conformation then is E = i<j ij B ij, where is unity when residues i and j form a native contact, and zero otherwise. Here, a contact formed between residues i and j may have different energies in different conformations, i.e., B ij may change from one conformation to 3

another (for the Gō model, one always has B ij = ε). In general, a contact formed between residues i and j will stabilize, to some extent, other contacts that residue i or j formed with other residues. On the contrary, its breakage may destabilize those contacts as well. Therefore, the introduction of the degree of order for a residue into the potential function reflects the cooperativity between the residues. Although the correlation distance is small, only one lattice unit, the many-body effects are obviously included in our model. Figure 1 shows such a collective effect. The interaction energies of contact A-B (or B-C) are different when the other contact is present or not present. Clearly, the energy of state I 3 is lower than the sum of that of states I 1 and I 2, indicating the interaction non-additivity. Each contact is stabilized by the other contact due to the collective effect. Note that in this paper our model is called Gō+ model to distinguish from the Gō model. Now let us present the Monte Carlo simulations on the thermodynamic and kinetic features for both models. The mean first passage time (MFPT), as a common measure of folding rate, is calculated by an average of the first passage time (FPT) over 1000 runs. Each run begins with a random conformation, and ends when the native state is reached for the first time. The FPT is the Monte Carlo steps (MCS) consumed in a run. Generally, as the temperature T decreases, the population of the native state, P N, increases from zero to about unity. The degree of sharpness of changes in P N, similar to the rapidity in Ref. [14], is a measure of the cooperativity of the folding reaction. Figure 2 shows the population P N and the specific heat C v versus temperature T for a 36-mer chain for both models. P N is defined as P N = e EN/T / EΩ(E)e E/T, where Ω(E) is the density of states for energy E, E N is the energy of the native state. Ω(E) is calculated with the Monte Carlo histogram method [15]. From Fig.2 we can see that the folding transition for our Gō+ model is much sharper than that of the Gō model, i.e., a sharper change in P N. There is also a single peak in C v curve, but it is narrower than that of the Gō model. For our Gō+ model, the maximum of C v occurs at a temperature nearly the midpoint temperature of transition with P N =1/2, i.e., the difference between these two temperatures is quite small. This is consistent with recent studies on naturally occuring proteins [14,16], implying 4

a good cooperativity of folding in this model. Differently, such a temperature difference is large for the Gō model (see Fig.2), indicating that the folding of the Gō model is much less cooperative than that of the Gō+ model. Since the sharpness is only a qualitative description for the transition, we further calculate the equilibrium energy distribution at the folding transition temperature, T f. Figure 3 shows such distributions for both models. Clearly our Gō+ model shows a good bimodal behavior, and the denatured-state energy is distributed in a narrow region [see Fig. 3(a)]. This means clearly a two-state folding and there is basically no intermediate states at equilibrium. Differently, for the Gō model as shown in Fig. 3(b), there are many intermediate states and the bimodal behavior is not so significant as that in Fig.3(a). Thus for the Gō model the folding is not of a two-state. This is in agreement with Chan and Kaya s argument [5]. In experiments, a well-established criterion for two-state folding is that the van t Hoff enthalpy H vh around the transition midpoint is equal, or very close, to the calorimetric enthalpy H cal of the entire transition. In this work, we calculate the ratio H vh / H cal as suggested in Ref. [5] (here, the definition of H vh / H cal is equal to (k 2 ) 2 in Ref. [5]), and list the results in Table I. From Table I, we can clearly see the difference between the Gō model and our Gō+ model. The Gō model, which is considered as a model with minimal energetic frustrations, does not meet the calorimetric two-state criterion and gives out the value of H vh / H cal far from 1. Nevertheless, our model satisfies the criterion quite well (for real proteins, the value of H vh / H cal is 0.96±0.03 [17]). This, again, implies the two-state folding and the good cooperativity of our Gō+ model. Physically, the high cooperativity of our model may result from the narrow distribution of the denatured states and the high population of the native state at the folding temperature (see also Figs.2 and 3). In our model, the energy spectrum relating to various conformations is redistributed, comparing with that of the Gō model, due to the collective effect between interactions. As a result, the energies of non-native conformations are moved to higher energy levels and a larger energy gap is left between the non-native conformations and the native one (for the two models, the energies of the native state are the same). The large 5

energy gap makes the native state paticularly stable, which is believed to be a necessary condition for cooperative folding [18]. This may be the physical origin of the two-state cooperativity. It can be further explained from the viewpoint of the free energy profile. For our Gō+ model, as shown in Fig.4, the free energy profiles have broad activation barriers. The broad activation barriers can account for the large movement of transition state caused by mutation or temperature changes, and are considered as a common feature of the twostate folding [19]. Our numerical results are surprisingly consistent with a phenomenological speculation for the existence of such a free energy profile in Ref. [19]. It should be noted that the broad activation barriers are consistent with the narrow distribution of the denatured states. Now let us make a comparison of the foldability based on the plots of the MFPT versus P N for both models. Note that we use P N instead of the commonly used temperature T in the horizontal axis in Fig.5. This is because that an identical condition should be taken for the comparison. In lattice simulations, the temperature has an arbitrary unit and also has no direct relationship with the real temperature. The comparison between two different models at the same temperature may make no sense. Nevertheless, at an identical condition with the same P N, the differences in the foldability can be well-defined. This is similar to other conditions used previously [20]. From Fig. 5, we can see that the MFPT for our Gō+ model shows a slow decrease as P N increases, it reaches a minimum at P N 0.93, and then it increases. For the Gō model, there is also a minimum but at P N 0.71. It is clearly that when the native state is stable (say, P N 0.9), our Gō+ model folds significantly fast, i.e., the MFPT is smaller with one or two orders of magnitude than that of the Gō model. Physically, this can be explained as follows. From Eq.(1) we can easily see that the energy gain of forming a contact is usually smaller for our Gō+ model than that for the Gō model. At high temperatures, entropic contribution is dominant to the free energy barrier, and the loss of entropy is always undercompensated by the energy gain, thus the Gō+ model folds slower for its smaller energy gain. Whereas at low temperatures, folding is nearly a downhill process, and the loss of entropy is always overcompensated by the energy gain. 6

Therefore, for the Gō+ model, it is easier to escape from kinetic traps, and the folding is faster. Finally, we note that for the two models the pathways of reaching the transition state from the denatured state are different. Due to the high cooperativity in our Gō+ model, a good core, the assembly of non-polar residues, is formed much earlier at low temperatures than that in the Gō model. Detailed kinetic results will be reported elsewhere. We also note that similar results are obtained for different chain sizes. In conclusion, our Gō+ model, with many-body interactions depending on the local structures included, exhibits a good two-state folding behavior. Our results suggest that the evolution of interactions during the folding plays an important role in the two-state cooperativity in protein folding. We give a possible way how the interactions evolve in the folding, which may capture some essential features of the two-state folding. We expect further study could provide new insights into the mechanism of protein folding. We thank H.S. Chan, A. Maritan and D. Thirumalai for useful suggestions. This work was supported by the Foundation of NNSF (No.19625409, and No.10074030). Email address: wangwei@nju.edu.cn 7

REFERENCES [1] K.A. Dill and H.S. Chan, Nat. Struct. Biol. 4, 10, (1997); D. Baker, Nature (London) 405, 39, (2000); W.A. Eaton et al., Annu. Rev. Biophys. Biomol. Struct. 29, 327, (2000); J.N. Onuchic, L.-S. Zaida, and P.G. Wolynes, Annu. Rev. Phys. Chem. 48, 525, (1997); C.M. Dobson and M. Karplus, Curr. Opin. Struct. Biol. 9, 92, (1999); J. Wang and W. Wang, Nat. Struct. Biol. 6, 1033, (1999). [2] J.M. Sorenson and T. Head-Gordon, Fold. Des. 3, 523, (1998). [3] S.E. Jackson, Fold. Des. 3, R81, (1998). [4] H.S. Chan, S. Bromberg, and K.A. Dill, Phil. Trans. R. Soc. Lond. B 348, 61, (1995); D.K. Klimov and D. Thirumalai, Fold. Des. 3, 127, (1998); [5] H.S. Chan, Proteins 40, 543, (2000); H. Kaya and H.S. Chan, Proteins 40, 637, (2000); H. Kaya and H.S. Chan, Phys. Rev. Lett. 85, 4823, (2000); [6] S. Miyazawa and R.L. Jernigan, Macromolecules 18, 534, (1985); A. Koliski, A. Godzik, and J. Skolnick, J. Chem. Phys. 98, 7420, (1993). [7] A.R. Fersht et al., Proc. Natl. Acad. Sci. USA 91, 10426, (1994); R.L. Baldwin, Nature (London) 369, 183, (1994). [8] K.A. Dill, J. Biol. Chem. 272, 701, (1997). [9] D.E. Otzen et al., Proc. Natl. Acad. Sci. USA 91, 10422, (1994); F. Chiti et al., Nat. Struct. Biol. 6, 1005, (1999); J.C. Martinez and L. Serrano, ibid. 6, 1010, (1999); D.S. Riddle et al., ibid. 6, 1016, (1999). [10] J.R. Banavar and A. Maritan, Proteins 42, 433, (2001). [11] C.J. Camacho and D. Thirumalai, Proc. Natl. Acad. Sci. USA 90, 6369, (1993). [12] S. Takada, Z. Luthey-Schulten, and P.G. Wolynes, J. Chem. Phys. 110, 11616, (1999). 8

[13] N. Gō, Annu. Rev. Biophys. Bioeng. 12, 183, (1983). [14] F. Cecconi et al., cond-matt/0101229. [15] A.M. Ferrenberg and R.H. Swendsen, Phys. Rev. Lett. 63, 1195, (1989); N.D. Socci and J.N. Onuchic, J. Chem. Phys. 103, 4732, (1995). [16] C. Micheletti et al., Phys. Rev. Lett. 82, 3372, (1999). [17] P.L. Privalov, Adv. Protein Chem. 33, 167, (1979). [18] A. Sali, E.I. Shakhnovich, and M. Karplus, Nature (London) 369, 248, (1994). [19] M. Oliveberg et al., J. Mol. Biol. 277, 933, (1998); D.E. Otzen et al., Biochemistry 38, 6499, (1999); M. Oliveberg, Acc. Chem. Res. 31, 765, (1998). [20] V.I. Abkevich, A.V. Gutin, and E.I. Shakhnovich, J. Mol. Biol. 252, 460, (1995); D.K. Klimov and D. Thirumalai, Proteins 26, 411, (1996). [21] J. Lee, Phys. Rev. Lett. 71, 211, (1993); M.H. Hao and H.A. Scheraga, J. Phys. Chem. 98, 4940, (1994). 9

Table I: The ratios of H vh / H cal for the Gō model and our Gō+ model, respectively. Ten sequences are calculated for each chain size. 10

FIG. 1: Schematic illustration of collective effect between two interactions. From a state I 0 with three unstructured residues, the chain can be settled in a state I 1 (or I 2 ) with a contact A-B (or B-C) and an equilibrium constant K 1 (or K 2 ). A state I 3 with two contacts A-B and B-C can be reached from state I 1 or I 2, but with different equilibrium constants K 2 γ or K 1 γ. In state I 3, each interaction is stronger by a factor γ due to the existence of the other contact. FIG. 2: Population P N and specific heat C v varying with the temperature T for a 36-mer chain. FIG.3: Theenergydistributionforthesame36-merusedinFig.2,using(a)Gō+potential and (b) Gō potential at respective folding transition temperature, T f. FIG. 4: The free energy profile F(E) = E TS(E) of our Gō+ model at different temperatures, where entropy S(E) is calculated by using entropy sampling Monte Carlo method [21]. Here U, N and TS denote the unfolded state, native state and transition state, respectively. Note that the free energy profile at high temperature is overall shifted so that the unfolded states are overlapped. FIG. 5: MFPT versus P N for a 36-mer chain. 11

A K 1 A B I 1 C K 2 γ B A B C I 0 C K 2 A B C K 1 γ I 3 I 2 Figure 1 By K. Fan, J. Wang and W. Wang in "Modeling two-state..."

1.0 0.8 300 P N 0.6 200 0.4 100 0.2 Go+ 0.0 0.2 Go 0.4 0.6 T 0.8 0 1.0 Cv Figure 2 By K. Fan, J. Wang, and W. Wang in "Modeling two-state..."

Population 0.3 0.2 0.1 0.0 0.3 0.2 0.1 (a) (b) T f =0.77 T f =0.72 0.0-40 -30-20 -10 0 E Figure 3 By K. Fan, J. Wang, and W. Wang in "Modeling two-state..."

F(E) -32-36 -40 T=0.85 TS TS N T=0.75 U N 0 10 20 - E 30 40 Figure 4 By K. Fan, J. Wang, and W. Wang in "Modeling two-state..."

MFPT (MCS) 10 10 10 9 10 8 10 7 10 6 Go+ Go 10 5 0.2 0.4 0.6 0.8 1.0 P N Figure 5 By K. Fan, J. Wang, and W. Wang in "Modeling two-state..."

Chain size 27-mer 36-mer 48-mer H vh / H cal Go model Go+ model 0.62±0.01 0.91±0.01 0.60±0.01 0.91±0.01 0.74±0.01 0.95±0.01