Molecular Determinants of Temperature Dependence of Protein

Aug 10, 2017 - Pressure is a well-known environmental stressor that can either stabilize or destabilize proteins. The volumetric change upon protein u...
0 downloads 14 Views 687KB Size
Subscriber access provided by UNIVERSITY OF ADELAIDE LIBRARIES

Article

Molecular Determinants of Temperature Dependence of Protein Volume Change upon Unfolding Calvin R. Chen, and George I Makhatadze J. Phys. Chem. B, Just Accepted Manuscript • DOI: 10.1021/acs.jpcb.7b05831 • Publication Date (Web): 10 Aug 2017 Downloaded from http://pubs.acs.org on August 22, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry B is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Molecular Determinants of Temperature Dependence of Protein Volume Change upon Unfolding Calvin R Chen and George I Makhatadze* Department of Biological Sciences and Center for Biotechnology and Interdisciplinary Studies, Rensselaer Polytechnic Institute, 110 8th Street, Troy, NY 12180 USA *Corresponding author: George I. Makhatadze, Center for Biotechnology and Interdisciplinary Studies, Rensselaer Polytechnic Institute, 110 8th Street, Troy, NY 12180 USA Phone: (518) 2764417 E-mail: [email protected]

Abstract Pressure is a well-known environmental stressor that can either stabilize or destabilize proteins. The volumetric change upon protein unfolding determines the effect of pressure on protein stability, where negative volume changes destabilize proteins at high pressures. High temperature often accompanies high pressure, for example in the ocean depths near hydrothermal vents or near faults, so it is important to study the effect of temperature on the volumetric change upon unfolding. We previously detailed the magnitude and sign of the molecular determinants of volumetric change, allowing us to quantitatively predict the volumetric change upon protein unfolding. Here we present a comprehensive analysis of the temperature dependence of the volumetric components of proteins, showing that hydration volume is the primary component that defines expansivities of the native and unfolded states and void volume only contributes slightly to the folded state expansivity.

1 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 26

Introduction Extremophilic organisms such as piezophiles and thermophiles can live under extreme pressures or temperatures 1, 2. The eukaryote Xenophyophores and other microbacteria have been discovered living in the Mariana Trench, where pressures reach up to 1100 atm 3. Hydrothermal vents are oceanic structures exposed to high pressures and temperatures, >150 atm and 400°C, that facilitate diverse communities such as chemosynthetic bacteria 4 and larger organisms such as giant tube worms 5. The presence of myriad organisms living in such extreme conditions suggests unique adaptations in their biological processes, body structure, and even cellular machinery (i.e. proteins) 6, as the high temperatures and pressures can unfold proteins from mesophiles and abolish their activity 7-9. Conversely, some of these extremophiles cannot survive when extracted from their extreme environments 10, hinting that their proteins require specific pressure and temperature ranges to remain folded. The application of pressure is known to perturb the thermodynamic equilibrium between the native and unfolded states. Specifically, the pressure stability of a protein is directly related to its volumetric change upon unfolding, ∆VTot: డ୼ீ ቀ డ௉ ቁ = ∆ܸ୘୭୲ = ܸ୘୭୲,୙ − ܸ୘୭୲,୒ 1. ்

Proteins that are capable of withstanding high pressures would have less negative changes in volume or even positive sign. The total volume of a protein can be calculated by partitioning of geometric components (VSE) and volume changes due to hydration (VHyd) i.e. interactions with the solvent. Geometric volume is the volume enclosed within the molecular surface as defined by Richards 11 and is also referred to as the solvent-excluded volume (VSE). The VSE of a protein is comprised of its van der Waals (VvdW) and void volume (VVoid) (Figure 1). The van der Waals volume is the volume occupied by protein atoms, while the void volume is the volume within the molecular surface that is not occupied by protein atoms, a result of the imperfect packing of spherical protein atoms. The total volume of the protein in solution, VTot, can be defined as the sum of three components: ܸ୘୭୲ = ܸୗ୉ + ܸୌ୷ୢ = ܸ୴ୢ୛ + ܸ୚୭୧ୢ + ܸୌ୷ୢ

2.

The volume changes upon protein unfolding are defined as the difference in total volumes of the unfolded, VTot,U, and native, VTot,N, states. The van der Waals volume of the native state is somewhat smaller than the unfolded state due to the extensive intramolecular hydrogen bonding in the native state 12. However, the ∆VvdW will not contribute to the total volume changes upon protein unfolding, ∆VTot, because protein-protein hydrogen bonds will reform with solvent upon protein unfolding. ∆ܸ୘୭୲ = ܸ୘୭୲,୙ − ܸ୘୭୲,୒ = ∆ܸ୚୭୧ୢ + ∆ܸୌ୷ୢ

3.

In our previous study, we had discussed the contributions of the individual components, ∆VvdW, ∆VVoid and ∆VHyd, to the total volume changes upon unfolding of proteins. We demonstrated that explicitly simulating the native and unfolded states, calculating ∆VVoid considering unfolded state void volume, and calculating ∆VHyd using non-polar and polar coefficients fit from empirical data can quantitatively predict experimentally measured ∆VTot 13.

2 ACS Paragon Plus Environment

Page 3 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

The change in volume upon unfolding is temperature dependent and increases positively with temperature. This positive temperature dependence of ∆VTot is expressed as a positive change in expansivity, Δߝ = ቀ different expansivities, ߝ = ቀ

ௗ୼௏౐౥౪

ௗ் ௗ௏౐౥౪ ௗ்

ቁ , and implies that the unfolded and native states possess ௉

ቁ . Therefore, a protein that has its melting temperature ௉

reduced by denaturant or pH will undergo a more negative ∆VTot than the same protein in the absence of denaturant and vice versa. However, the underlying cause of this difference has heretofore been conflicted, with both internal voids 14 or hydration 15 being implicated as major contributors. Since a steeper change in expansivity leads to more positive ∆VTot (i.e. pressure stability) at higher temperatures, proteins evolving to have a larger difference in unfolded and native expansivity is one possible way to maintain stability in high pressure and high temperature environments such as near deep sea vents. Therefore, understanding the molecular determinants behind ε and ∆ε is critical for understanding the possible mechanisms used by extremophilic proteins to remain folded in extreme conditions. We previously developed a structural thermodynamic model that computes the individual molecular determinants of volumes of native and unfolded states. In this work, we expand this formalism to compute the temperature dependence of the individual volumetric components, VVoid and VHyd, of the native and unfolded states and their contributions to the temperature dependence of total volume changes upon unfolding of proteins.

Methods Computational Modeling of Native and Unfolded Ensembles The native state of a protein is not static, but a dynamic ensemble of structures. Therefore, ensemble sampling is necessary to determine the volumetric properties of the native state. Considering that molecular dynamics simulation is a well-established method to provide information about protein conformational fluctuations, we have used this method to model the fluctuations of native proteins 16. As a test case for all calculations we have selected 8 proteins that were studied experimentally before: bovine pancreatic trypsin inhibitor (BPTI), eglin C (Egl), bovine ubiquitin (UBQ), human acylphosphatase (ACP), ribonuclease A (RNase), hen egg white lysozyme (Lyz), staphylococcal nuclease (SNase), and staphylococcal nuclease ∆PHS variant (SNase ∆PHS). For each protein, all-atom explicit solvent MD simulation were carried out using identical settings as our previous study 13, with the change of simulation length which here was extended to 400 nsec. In brief, we used the molecular dynamics software package GROMACS 4.6.3 17 to simulate all structures using the CHARMM22 force field with CMAP corrections, which is colloquially referred to as the CHARMM27 force field in GROMACS, and the TIP3P water model 18. Simulation conditions were 0.1 molar excess NaCl, 1 bar of pressure, and temperatures ranging from 300 to 400 K in 20 K increments. All protein simulations underwent 1,000 steps of energy minimization, 200 psec of constant volume equilibration, 200 psec of constant pressure equilibration, and 400 nsec of production simulation. A native ensemble of 400 structures was extracted from the production trajectory (1 structure/ns). Equilibration criteria were stable all-atom RMSD with respect to the crystal structure (0 - 5 Å) and stable 3 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

values (drift 0.5%) of solvent-excluded volume throughout the production trajectory. Simulations were started from the three-dimensional coordinates as provided in the PDB file, BPTI - PDB:6PTI, Egl - PDB:1EGL, Ubq - PDB:1UBQ, ACP PDB:2K7J, RNase - PDB:7RSA, Lyz - PDB:4LZT, SNase - PDB:1EY0, and SNase ∆PHS – PDB:3BDC. All proteins remained folded throughout the simulation at all temperatures (Supplementary Figures S1 and S2). In addition, volume calculations were performed directly (i.e. omitting MD simulations) on the crystal structures of RNase (PDB: 1-9RAT) and Lyz (5-6LYT) that were solved at different temperatures. Prior to the volume calculations, the structures were energy minimized using the CHARMM27 force field and implicit solvent (Generalized Born surface area, GBSA, model) with protein and solvent dielectric constants of 80 in GROMACS 4.6.3 to incorporate all hydrogen atoms. Explicitly modeling the unfolded state ensemble is important to draw conclusions about the volumetric properties of the unfolded state 13. There are multiple computational ways to generate the unfolded state ensemble, including statistical coil model of the unfolded state (SC) 19 , trajectory directed ensemble sampling (TraDES) 20, and flexible-meccano (FM) 21. We have shown previously, TraDES generated unfolded state ensemble shows remarkable agreement to experimental values of Rg as a function of protein size 13, 22. Typically TraDES generated unfolded state ensemble consists of 1,000 structures, that are generated with the all-coil sampling flag (-c T) to remove all secondary structure propensity. Prior to the volume calculations, the structures are energy minimized in implicit solvent using the Generalized Born surface area (GBSA) model with protein and solvent dielectric constants of 80 in GROMACS 4.6.3. This step also explicitly incorporates all hydrogen atoms. The unfolded ensemble of three proteins, BPTI, RNase, and Lyz were simulated with their disulfide bonds intact using the structure-based model (SBM) generator SMOG 23-25. The crystal structure files of BPTI (PDB:6PTI), Ribonuclease (PDB:7RSA), and Lysozyme (PDB:4LZT) were appended with the sulfur atom numbers participating in the disulfide bond. The modified PDB files were then input into SMOG and processed with the all-atom and shadow contact models. The resulting output were GROMACS-compatible coordinate and topology files. The output topology files were stripped of all pairwise interactions, leaving only the Lennard-Jones, bond, angle, dihedral, and improper contributions to the energy function. Binary run input files (tpr) files were generated from the coordinate and modified topology files. The SBM simulations were run at 150 K for 108 steps with a time step of 0.2 psec. The simulation conditions were no solvent, no periodic boundaries, no charges, and a short-range neighbor list and vdW cut-off of 1.2 nm. Equilibration of radius of gyration was seen after 500 psec, so extraction of PDB structures started from 500 psec in 10 psec intervals, resulting in a total of 1,950 unfolded structures per disulfide bond-containing protein. To better understand the effects of temperature on the volumetric properties of the unfolded state ensemble, we have also performed all-atom explicit solvent MD simulations of an intrinsically disordered peptide with the sequence GIHKQKEKSRLQGGVLVNEILNHMKRATQIPSYKKLIMY. Simulations were done using enhanced sampling approach using replica exchange molecular dynamics (REMD). 48 replicas spanning the temperature range from 280 K to 450 K were used, and simulation was carried out for 1.1 µsec per replica (~53 µsec total) using the CHARMM27 force field and TIP3P water model. The volumetric

4 ACS Paragon Plus Environment

Page 4 of 26

Page 5 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

properties were calculated using structural ensembles of the peptide at different temperatures from the last 300 nsec of simulations. Model Compounds for Calculating Volume of Hydration The experimentally measured partial molar volumes in aqueous solution, Vϕ,aq, of 106 model compounds relevant to proteins (alkanes, aromatic compounds, alcohols, diols, amines, amides, polyethylene glycols, carboxylic acids, and acetals) were taken from 26-38. Partial molar volumes in aqueous solution for these model compounds were experimentally measured at different temperatures, typically ranging from 273K to 328K (0 to 55°C). For each of these model compound PDB structures were generated using the CORINA webserver 39. Additionally, we included into our volumetric analysis more complex model compounds such as oligopeptides (3-5 residues), N-acetyl amino acid amides, and N-acetyl amino acids with neutralized carboxyl termini that represent protein components 40, 41. For these compounds, structures were generated using the same protocol as for the native proteins. The hydration volume is calculated as a difference between partial molar volume of a compound in aqueous solution, Vϕ,aq, and corresponding solvent-excluded volume, VSE. MSMS software package 42 was used to analytically calculate MSA per atom for all model compound structures. For each structure MSA was broken down into carbon, nitrogen, oxygen, and sulfur surface areas. Hydrogen surface areas were combined with the surface area of their parent heavy atom. Carbon and sulfur molecular surface areas were combined into non-polar MSA (MSANP). Nitrogen and oxygen molecular surface areas were combined into polar MSA (MSAPol). ProteinVolume Calculations The ProteinVolume (http://gmlab.bio.rpi.edu) software package has been described by us previously 43. The input is a PDB-formatted coordinates file, which is used to generate the accessible surface using a flood-fill algorithm operating in the polar coordinate space. The Bondi radii set 44 is used for all-atom representation. The enclosed volume within the accessible surface is flood-filled with probes of variable length, ranging from 0.08 Å to 0.02 Å, depending on the complexity of the volume. The flood-fill algorithm stops at 1.4 Å from the accessible surface. The existing volume probes describe the volume enclosed within the molecular surface, or solvent-excluded volume (VSE). Probes that lie within the volume of a protein atom are considered to be part of the van der Waal’s volume (VvdW). A probe which lies on top of a van der Waals boundary is stochastically accepted with the acceptance probability based on its magnitude of overlap with the atom. Void volume, VVoid, is calculated as the difference between the solvent-excluded volume and the van der Waals volume. Structural Analysis Methods Contact maps, Rg, and number of hydrogen bonds were analyzed using GROMACS 4.6.3 utilities g_mdmat, g_gyrate, and g_hbond. Secondary structure was calculated using DSSP 45. The number of nearby atomic contacts was calculated using shadow contact map (SCM) 24 using a flat cutoff distance of 4 Å.

5 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 26

Results and Discussion Calculating Volume of Hydration and its Change upon Unfolding When a protein molecule is solvated, there are changes in the surrounding solvent volume, relative to the bulk, due to interactions with the protein surface. This volume change is referred to as hydration volume (VHyd). In our previous work 13, volumetric data at 25°C of 150 model compounds was used to formulate a hydration volume model that scales linearly with nonpolar and polar molecular surface area: ܸୌ୷ୢ = ܽ + ݇୒୔ ∙ ‫ ୔୒ܣܵܯ‬+ ݇୔୭୪ ∙ ‫୔ܣܵܯ‬୭୪ 4. The volume of a solute changes with temperature, therefore we wanted to understand the degree of volumetric expansion that was originating from VHyd. To this end, we analyzed the experimentally measured partial molar volumes for over 100 model solutes in aqueous solutions at temperatures ranging from 273-363K (0 - 90°C) with the majority of data obtained in the 280340 K range 26-38, 40, 41. These compounds contain functional groups relevant to proteins, i.e. alkanes, aromatic compounds, alcohols, diols, amines, amides, polyethylene glycols, carboxylic acids, acetals, oligopeptides, and N-acetyl amino acid amides. Hydration volume, VHyd, is equal to the volume change upon transfer of a solute from gas phase into water, i.e. the partial volume of a solute in water minus the geometric volume of this solute. A linear regression fit was performed on the hydration volumes per model compound across the temperature range of 280340 K. The temperature dependence of the hydration coefficients kNP and kPol is shown in Figure 2, while the quality of the fit is shown in Supplementary Figure S3. The kNP coefficient has larger value at 280 K and shows relatively steep linear dependence on temperature in the experimental range (280-340 K). In contrast, kPol coefficient at 280 K has smaller value than kNP coefficient and shows a more moderate linear dependence on temperature (Figure 2). The small value for the polar coefficient is due to polar groups hydrogen bonding with water, while the large value of the kNP coefficient suggests that water molecules move away from non-polar groups due to the hydrophobic effect 7, 46, 47. The increase in kNP and kPol with temperature can be interpreted as the increase in equilibrium distance between solvent and the protein surface due to increased energy. The kNP coefficient increases more steeply with temperature compared to kPol because polar surfaces hydrogen bond with solvent and reduce the expansion of the hydration volume around polar surfaces with temperature. With the non-polar and polar hydration coefficients known, the contribution of hydration to the volume of native and unfolded states, and to the changes in hydration volume upon unfolding, ∆VHyd, can be calculated as using the following equations 5-7. ܸୌ୷ୢ,୒ = ݇୒୔ ∙ ‫୔୒ܣܵܯ‬,୒ + ݇୔୭୪ ∙ ‫୔ܣܵܯ‬୭୪,୒ ܸୌ୷ୢ,୙ = ݇୒୔ ∙ ‫୔୒ܣܵܯ‬,୙ + ݇୔୭୪ ∙ ‫୔ܣܵܯ‬୭୪,୙ ∆ܸୌ୷ୢ = ݇୒୔ ∙ ∆‫ ୔୒ܣܵܯ‬+ ݇୔୭୪ ∙ ∆‫୔ܣܵܯ‬୭୪

5. 6. 7.

where ∆MSANP=MSANP,U-MSANP,N and ∆MSAPol=MSAPol,U-MSAPol,N are the differences in the nonpolar and polar surface areas of the unfolded and native states, respectively. This formalism assumes that the proportionality coefficients kNP and kPol derived from model compounds (i.e. equation 4) can be extrapolated to much larger surfaces. Expansivity of the Native State 6 ACS Paragon Plus Environment

Page 7 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

The partial volume of native proteins in aqueous solution increases with increase in temperature 14, 28. It has been argued that this volumetric expansion is due to the increase in void volume as internal fluctuations in the protein increase due to the increase in thermal motion. To test this hypothesis, we have performed all-atom explicit solvent simulations of the native state ensembles for eight different proteins at six different temperatures ranging from 300 K to 400 K in 20 K increments (Methods section for simulation details and Supplementary Figure S1). Analysis of these structural ensembles allowed us to compute the solvent-excluded, VSE,N, and void, VVoid,N, volumes at these temperatures. Figure 3 shows the temperature dependencies of VSE,N, and VVoid,N. From the figure, it is evident that both VSE,N and VVoid,N functions do not show significant changes at 400 K compared to 300 K. Any slight change in VSE,N is mirrored in VVoid,N, indicating that, as expected, VvdW is temperature independent. The relative changes in VSE,N and VVoid,N per 100 K for the 8 proteins calculated from molecular dynamics simulations are 0.5±0.4% (max 1.1%, min 0.1%) and 2.1±1.5% (max 4.3%, min 0.5%) respectively. The absolute changes for the 8 proteins obtained from molecular dynamics simulations are 80±70 Å3 (max 186 Å3, min 15 Å3) per 100 K for VSE,N and VVoid,N. To understand why VSE,N and VVoid,N increases with temperature, we analyzed the number of contacts within 4 Å as a function of temperature (Figure 4), which revealed a strong negative linear dependence of nearby contacts on temperature. Considering that VvdW is temperature independent, the small increase in VVoid,N and thus in VSE,N is likely caused by the reduction of intramolecular contacts which results in the expansion of void volume in between the previously contacting but non-covalently-bonded protein atoms. This slight increase in VSE and VVoid with temperature has also been observed in crystal structures of RNase 48 and Lyz 49 solved at temperatures ranging from 98K to 298K. Ribonuclease displayed a 63 Å3 (0.4% and 1.8%) increase in VSE, and VVoid per 100K (see Supplementary Figure S4), which is very similar to our value of 145 Å3 (0.9%) increase in VSE per 100K calculated from our MD simulation. Likewise, increases in volumes for Lyz are 120 Å3 (0.7% and 2.1%) per 100 K for VSE, and VVoid, respectively, similar to 187 Å3 (1.1%) obtained from MD simulations. This similarity of the computed and experimental volume changes with temperature provides additional support that molecular dynamics simulations can provide realistic estimates for void volume expansion within a protein across a 100K temperature range. The contribution of hydration to the volume of the native state, VHyd,N, taken into account using equation 5, shows a significant temperature dependence and thus appears to be a major determinant for the temperature dependence of the volume of the native proteins, i.e. VTot,N (Figure 3). Average increase in VHyd,N per 100K for the 8 proteins is 530±140 Å3 (min 330 Å3, max 690 Å3). The temperature dependence of hydration volume is dependent on MSANP,N, MSAPol,N, kNP, and kPol (Equation 5). To understand which of these components are responsible for the thermal expansion of hydration volume, we analyzed their magnitudes from 300 - 400K. As we have shown above, the kNP and kPol coefficients demonstrate positive temperature dependence, with kNP increasing much more quickly with temperature. There is also a slight increase in MSANP,N and MSAPol,N with temperature (Supplementary Figure S5) but their increase across 300 - 400K is only ~7% whereas kNP increases by 28% (Figure 2 and Supplementary Table S1). Such minor increases in surface area with increasing temperature is expected given the overall nativeness of the MD ensembles (see Supplementary Figure 1). Since the relative increase in kNP and kPol with temperature is much larger in magnitude than the increase in MSANP,N and MSAPol,N, the

7 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

temperature dependence of the hydration volume is primarily due to the increasing kNP and kPol coefficients with temperature. To summarize, the folded proteins display a positive expansivity ε= (∂V/∂T)P which has contributions from temperature dependencies of both VVoid,N or VHyd,N. Our analysis also shows that the thermal expansion originating from the hydration volume is primarily responsible for the thermal expansion of total volume, contributing ~85% to the expansivity of the native state. Expansivity of the Unfolded State As described in the previous section, the expansivity of the native state is primarily a function of hydration volume (85-90%) with minor (10-15%) contributions from void volume. The compact three-dimensional structure of the native state maintained by long-range interactions allows void volume to undergo slight expansion with temperature due to weakening of intramolecular interactions. However, due to the lack of the long-range specific contacts in the unfolded state, increase in temperature can only affect the overall dimensions of the polypeptide chain, leading to its expansion as characterized by, for example, increase in radius of gyration (Rg) as a function of temperature. The increase in Rg within the unfolded state ensemble should not lead to significant changes in the void or solvent excluded volume and they can be assumed to be temperature-independent. This assumption was tested using a 39 amino acid long intrinsically-disordered peptide that due to its small size can be simulated using extended (>1 microsec) all-atom explicit solvent molecular dynamics simulations (see Methods for details). While certain forcefields can overcompact the unfolded state of intrinsically disordered proteins, we previously demonstrated the insensitivity of VSE to radius of gyration of the unfolded state ensemble13. The average Rg values fall within the range expected for the unfolded polypeptide of this size and increase gradually as the temperature increases (see Supplementary Figure S6). However, the corresponding void and solvent excluded volumes at 300 K and 400 K are the same within the error of calculation (Figure 5). Moreover, the corresponding MSANP, and MSAPol are also independent of temperature (see Supplementary Figure S6). This supports the validity of the assumption that the unfolded state ensemble VSE, VVoid, MSANP, and MSAPol are independent of temperature. This establishes that the VSE and VVoid of the unfolded state for the 8 proteins considered here are independent of temperature. Thus, VHyd is the only volumetric component that increases with temperature in the unfolded state and can calculated by equation 6, using the values of MSANP and MSAPol and the respective kNP and kPol coefficients from 300 - 400K. Consequently, the expansivity of the unfolded state is exclusively defined by the expansivity of its hydration volume (Figure 6). Change in expansivity (∆ε) upon protein unfolding and comparison with experiment Most proteins undergo a negative volumetric change upon unfolding at lower temperatures around 300K 50, 51, indicating that the volume of the unfolded state is smaller than the volume of the native state. However, at higher temperatures around 360K, the volumetric change upon unfolding becomes less negative and sometimes even positive 50, indicating that the unfolded state volume has expanded to become larger than the native state due to the higher expansivity of the unfolded state compared to the native state 52, 53. This experimental observation is recapitulated in our computational analysis. Indeed, the temperature dependence of the unfolded state computed for 8 proteins studied here are steeper than the temperature dependence of the native state (compare the slopes of VTot vs temperature in Figures 3 and 6). 8 ACS Paragon Plus Environment

Page 8 of 26

Page 9 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

The changes in expansivity upon unfolding can be calculated as: Δߝ = ߝ୙ − ߝ୒ = ቀ

ௗ∆௏౐౥౪ ௗ்

ቁ =ቀ ௉

ௗ∆௏౒౥౟ౚ ௗ்

ቁ +ቀ ௉

ௗ∆௏ౄ౯ౚ ௗ்

ቁ = Δߝ௏௢௜ௗ + Δߝு௬ௗ ௉

8.

This allows for the analysis of the contributions of the individual components, i.e. changes in expansivity of the void volume, ∆εVoid, and the hydration volume, ∆εHyd, to the overall ∆ε. We have established in the previous sections that: 1) hydration volume is the primary contributor to expansivity of both native and unfolded states (Figures 3 and 6); 2) hydration volume is directly proportional to polar and non-polar MSA (equation 7); and 3) the unfolded state has a much higher polar and non-polar MSA than the native state (Supplementary Figure S5). Therefore, the unfolded state will always have a higher expansivity than the native state because the unfolded state has more hydration volume, the primary component of the thermal expansivity of both states. Although the void volume expansivity of the native state contributes slightly to the native state expansivity and has no contribution to unfolded state expansivity, its magnitude is much smaller than the difference in hydration volume expansivity between the unfolded and native states and therefore the unfolded state expansivity remains higher. The contribution from changes in void volume with temperature, ∆εVoid, to the overall expansivity is negative, i.e. ∆VVoid decreases with increase in temperature (Figure 7). The contribution from changes in hydration volume with temperature, ∆εHyd, to the overall expansivity is positive and larger in absolute magnitude than the negative ∆εVoid (Figure 7). This leads to the positive overall change in expansivity, ∆ε, for all proteins studied here (Figure 7) which has been observed experimentally 50, 54. Figure 8 compares experimental and computed values of overall changes in expansivity. It is evident that there is good qualitative agreement between the magnitude of calculated and experimental ∆ε values with a correlation R2 = 0.76. A possible source of the observed systematic overestimation ∆ε is the assumption of constant VVoid,U across all temperatures. There is some indication from our all-atom explicit solvent molecular dynamics simulation of the unfolded state ensemble that VVoid,U might be decreasing with increasing temperature (Figure 5), however this trend is masked by the accuracy in computing VVoid,U. If this trend is real, it will lead to lower values of VTot,U at higher temperatures and thus lower ∆ε. Nevertheless, while there appears to be a systematic overestimation of ∆ε, the model can predict relative differences in ∆ε between proteins with remarkable accuracy. Concluding Remarks We have analyzed the contributing factors to the expansivity of native and unfolded states, which revealed that hydration volume is the primary contributor to the temperature dependence of volume, with void volume playing a minor role in expansivity. These findings explain why the native state has a lower expansivity than the unfolded state, i.e. due to the lower molecular surface area in the native state and consequently lower hydration volume in the native state compared to the unfolded state. Our model can also reasonably well predict the relative differences in ∆V 13 and ∆ε (this work). This will allow for the computational prediction of expected differences in ∆V and ∆ε for unfolding of homologous proteins and in the longer run, optimization of proteins with more positive changes in volume at higher temperature. The possible avenues for increasing ∆V and ∆ε are 1) burying more non-polar surface area within the protein core via polar to non-polar mutations or small to large non-polar mutations and 2) improving the packing of the core by 9 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 26

mutating large bulky residues to smaller non-polar residues. It is important to note that the two methods seemingly counteract each other, as mutating a small non-polar residue to a larger one may increase the native state expansivity through the disruption of packing and reduction of intramolecular interactions, which increases the native state void volume. Likewise, substituting a large non-polar residue within the core with a smaller non-polar residue may decrease the expansivity of the unfolded state to a greater magnitude, due to the reduction in unfolded state non-polar surface area, than it decreases the expansivity of the native state through better packing. Therefore, the general finding that hydration volume contributes the majority share to thermal expansivity of proteins hints that the first method of modulating the exposed surface area upon unfolding will produce greater results for engineering changes in ∆ε, with the important caveat that the mutations do not significantly destabilize the protein core.

10 ACS Paragon Plus Environment

Page 11 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Supporting Information Parameters characterizing native state ensembles of 8 proteins (Figures S1 & S2). Quality of fit for hydration (Figure S3). Temperature dependence of volumes from x-ray (Figure S4). Changes in the molecular surface areas of 8 proteins as a function of temperature (Figure S5). Parameters characterizing the unfolded state ensemble of a model peptide (Figure S6). Coefficients of volume of hydration as a function of temperature (Table S1). Volumes of hydration as a function of temperature for model compounds (Table S2). This material is available free of charge via the Internet at http://pubs.acs.org.

Acknowledgements Supported by a grant from the US National Science Foundation CHE/CLP-1145407 (to GIM). This work used the Extreme Science and Engineering Discovery Environment (XSEDE), which is supported by National Science Foundation grant number ACI-1053575, and Center for Computational Innovations at Rensselaer Polytechnic Institute.

11 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 26

REFERENCES (1) Horikoshi, K. Barophiles: deep-sea microorganisms adapted to an extreme environment. Curr. Opin. Microbiol. 1998, 1, 291-295. (2) Lauro, F. M.; Bartlett, D. H. Prokaryotic lifestyles in deep sea habitats. Extremophiles 2008, 12, 15-25. (3) Gallo, N. D.; Cameron, J.; Hardy, K.; Fryer, P.; Bartlett, D. H.; Levin, L. A. Submersibleand lander-observed community patterns in the Mariana and New Britain trenches: Influence of productivity and depth on epibenthic and scavenging communities. Deep Sea Res Part 1 Oceanogr Res Pap 2015, 99, 119-133. (4) Cavanaugh, C. M. Symbiotic chemoautotrophic bacteria in marine invertebrates from sulphide-rich habitats. Nature 1983, 302, 58-61. (5) Black, M. B.; Halanych, K. M.; Maas, P. A. Y.; Hoeh, W. R.; Hashimoto, J.; Desbruyères, D.; Lutz, R. A.; Vrijenhoek, R. C. Molecular systematics of vestimentiferan tubeworms from hydrothermal vents and cold-water seeps. Mar. Biol. 1997, 130, 141-149. (6) Huang, Q.; Tran, K. N.; Rodgers, J. M.; Bartlett, D. H.; Hemley, R. J.; Ichiye, T. A molecular perspective on the limits of life: Enzymes under pressure. Condensed Matter Physics 2016, 19, 1-16. (7) Hummer, G.; Garde, S.; García, a. E.; Paulaitis, M. E.; Pratt, L. R. The pressure dependence of hydrophobic interactions is consistent with the observed pressure denaturation of proteins. Proc. Natl. Acad. Sci. U. S. A. 1998, 95, 1552-1555. (8) Hei, D. J.; Clark, D. S. Pressure stabilization of proteins from extreme thermophiles. Appl. Environ. Microbiol. 1994, 60, 932-939. (9) Bartlett, D. H. Pressure effects on in vivo microbial processes. Biochim. Biophys. Acta 2002, 1595, 367-381. (10) Zeng, X.; Birrien, J.-L.; Fouquet, Y.; Cherkashov, G.; Jebbar, M.; Querellou, J.; Oger, P.; Cambon-Bonavita, M.-A.; Xiao, X.; Prieur, D. Pyrococcus CH1, an obligate piezophilic hyperthermophile: extending the upper pressure-temperature limits for life. The ISME journal 2009, 3, 873-876. (11) Richards, F. M. Areas, volumes, packing and protein structure. Annu. Rev. Biophys. Bioeng. 1977, 6, 151-176. (12) Schell, D.; Tsai, J.; Scholtz, J. M.; Pace, C. N. Hydrogen bonding increases packing density in the protein interior. Proteins 2006, 63, 278-282. (13) Chen, C. R.; Makhatadze, G. I. Molecular determinant of the effects of hydrostatic pressure on protein folding stability. Nat Commun. 2017, 8, 14561-14561. (14) Dellarole, M.; Kobayashi, K.; Rouget, J. B.; Caro, J. A.; Roche, J.; Islam, M. M.; GarciaMoreno, E. B.; Kuroda, Y.; Royer, C. A. Probing the physical determinants of thermal expansion of folded proteins. J. Phys. Chem. B 2013, 117, 12742-12749. (15) Royer, C. A. Insights into the role of hydration in protein structure and stability obtained through hydrostatic pressure studies. Braz. J. Med. Biol. Res. 2005, 38, 1167-1173. (16) Daggett, V.; Levitt, M. Realistic simulations of native-protein dynamics in solution and beyond. Annu. Rev. Biophys. Biomol. Struct. 1993, 22, 353-380. (17) Pronk, S.; Páll, S.; Schulz, R.; Larsson, P.; Bjelkmar, P.; Apostolov, R.; Shirts, M. R.; Smith, J. C.; Kasson, P. M.; van der Spoel, D., et al. GROMACS 4.5: a high-throughput and highly parallel open source molecular simulation toolkit. Bioinformatics 2013, 29, 845-854.

12 ACS Paragon Plus Environment

Page 13 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(18) Brooks, B. R.; Brooks, C. L.; Mackerell, A. D.; Nilsson, L.; Petrella, R. J.; Roux, B.; Won, Y.; Archontis, G.; Bartels, C.; Boresch, S., et al. CHARMM: The biomolecular simulation program. J. Comput. Chem. 2009, 30, 1545-1614. (19) Jha, A. K.; Colubri, A.; Freed, K. F.; Sosnick, T. R. Statistical coil model of the unfolded state: resolving the reconciliation problem. Proc. Natl. Acad. Sci. U. S. A. 2005, 102, 1309913104. (20) Feldman, H. J.; Hogue, C. W. Probabilistic sampling of protein conformations: new hope for brute force? Proteins 2002, 46, 8-23. (21) Ozenne, V.; Bauer, F.; Salmon, L.; Huang, J.-R.; Jensen, M. R.; Segard, S.; Bernadó, P.; Charavay, C.; Blackledge, M. Flexible-meccano: a tool for the generation of explicit ensemble descriptions of intrinsically disordered proteins and their associated experimental observables. Bioinformatics 2012, 28, 1463-1470. (22) Fitzkee, N. C.; Rose, G. D. Reassessing random-coil statistics in unfolded proteins. Proc. Natl. Acad. Sci. U. S. A. 2004, 101, 12497-12502. (23) Whitford, P. C.; Noel, J. K.; Gosavi, S.; Schug, A.; Sanbonmatsu, K. Y.; Onuchic, J. N. An all-atom structure-based potential for proteins: bridging minimal models with all-atom empirical forcefields. Proteins 2009, 75, 430-441. (24) Noel, J. K.; Whitford, P. C.; Onuchic, J. N. The shadow map: A general contact definition for capturing the dynamics of biomolecular folding and function. J. Phys. Chem. B 2012, 116, 8692-8702. (25) Noel, J. K.; Levi, M.; Raghunathan, M.; Lammert, H.; Hayes, R. L.; Onuchic, J. N.; Whitford, P. C. SMOG 2: A versatile software package for generating structure-based models. PLoS Comput. Biol. 2016, 12, e1004794. (26) Alexander, D. Apparent molar volumes of alcohols in dilute aqueous solution. J. Chem. Eng. Data 1959, 4, 252-254. (27) Friedman, M. E.; Scheraga, H. A. Volume changes in hydrocarbon-water systems. Partial molal volumes of alcohol-water solutions 1. J. Phys. Chem. 1965, 69, 3795-3800. (28) Makhatadze, G. I.; Medvedkin, V. N.; Privalov, P. L. Partial molar volumes of polypeptides and their constituent groups in aqueous solution over a broad temperature range. Biopolymers 1990. (29) Cabani, S.; Conti, G.; Lepori, L. Volumetric properties of aqueous solutions of organic compounds. III. Aliphatic secondary alcohols, cyclic alcohols, primary, secondary, and tertiary amines. J. Phys. Chem. 1974, c, 2-6. (30) Harada, S.; Nakajima, T.; Komatsu, T.; Nakagawa, T. Apparent molal volumes and adiabatic compressibilities of ethylene glycol derivatives in water at 5, 25, and 45°C. J. Solution Chem. 1978, 7, 463-474. (31) Nakajima, T.; Komatsu, T.; Nakagawa, T. Apparent molal volumes and adiabatic compressibilities of n-alkanols and a,w-alkane diols in dilute aqueous solutions at 5, 25 and 45C. I. Apparent molal volumes. Bulentin of the Chemical Society of Japan 1975, 48, 783-787. (32) Pagé, M.; Huot, J.-Y.; Jolicoeur, C. A thermodynamic investigation of aqueous 2methoxyethanol. J. Chem. Thermodyn. 1993, 25, 139-154. (33) Pal, A.; Singh, Y. P. Excess molar volumes and apparent molar volumes of some amide + water systems at 303.15 and 308.15 K. J. Chem. Eng. Data 1995, 40, 818-822. (34) Krakowiak, J.; Wawer, J.; Panuszko, A. Densimetric and ultrasonic characterization of urea and its derivatives in water. J. Chem. Thermodyn. 2013, 58, 211-220.

13 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 26

(35) Shvedov, D.; Tremaine, P. R. Thermodynamic properties of aqueous dimethylamine and dimethylammonium chloride at temperatures from 283 K to 523 K: Apparent molar volumes, heat capacities, and temperature dependence of ionization. J. Solution Chem. 1997, 26, 11131143. (36) Roux, G.; Perron, G.; Desnoyers, J. E. The heat capacities and volumes of some low molecular weight amides, ketones, esters, and ethers in water over the whole solubility range. Can. J. Chem. 1978, 56, 2808-2814. (37) Cibulka, I. Partial molar volumes and partial molar isentropic compressions of four poly(ethylene glycols) at infinite dilution in water at temperatures T= (278 to 343) K and atmospheric pressure. J. Chem. Eng. Data 2016, 61, 748-759. (38) Hnědkovský, L.; Cibulka, I. Partial molar volumes and partial molar isentropic compressions of selected alkane-α,ω-diols at infinite dilution in water at temperatures T = (278 to 318) K and atmospheric pressure. J. Chem. Eng. Data 2013, 58, 1724-1734. (39) Sadowski, J.; Gasteiger, J.; Klebe, G. Comparison of automatic three-dimensional model builders using 639 X-ray structures. J. Chem. Inf. Model. 1994, 34, 1000-1008. (40) Lee, S.; Tikhomirova, A.; Shalvardjian, N.; Chalikian, T. V. Partial molar volumes and adiabatic compressibilities of unfolded protein states. Biophys. Chem. 2008, 134, 185-199. (41) Hedwig, G. R.; Hinz, H.-J. Group additivity schemes for the calculation of the partial molar heat capacities and volumes of unfolded proteins in aqueous solution. Biophys. Chem. 2003, 100, 239-260. (42) Sanner, M. F.; Olson, A. J.; Spehner, J. C. Reduced surface: an efficient way to compute molecular surfaces. Biopolymers 1996, 38, 305-320. (43) Chen, C. R.; Makhatadze, G. I. ProteinVolume: calculating molecular van der Waals and void volumes in proteins. BMC Bioinformatics 2015, 16, 1-6. (44) Bondi, A. van der Waals volumes and radii. J. Phys. Chem. 1964, 68, 441-451. (45) Kabsch, W.; Sander, C. Dictionary of protein secondary structure: Pattern recognition of hydrogen-bonded and geometrical features. Biopolymers 1983, 22, 2577-2637. (46) Giovambattista, N.; Lopez, C. F.; Rossky, P. J.; Debenedetti, P. G. Hydrophobicity of protein surfaces: Separating geometry from chemistry. Proc. Natl. Acad. Sci. U. S. A. 2008, 105, 2274-2279. (47) Dadarlat, V. M.; Post, C. B. Adhesive-cohesive model for protein compressibility: An alternative perspective on stability. Proc. Natl. Acad. Sci. U. S. A. 2003, 100, 14778-14783. (48) Tilton, R. F.; Dewan, J. C.; Petsko, G. A. Effects of temperature on protein structure and dynamics: X-ray crystallographic studies of the protein ribonuclease-A at nine different temperatures from 98 to 320 K. Biochemistry 1992, 31, 2469-2481. (49) Young, A. C. M.; Tilton, R. F.; Dewan, J. C. Thermal expansion of hen egg-white lysozyme. Comparison of the 1.9 A resolution structures of the tetragonal form of the enzyme at 100 K and 298 K. J. Mol. Biol. 1994, 235, 302-317. (50) Schweiker, K. L.; Fitz, V. W.; Makhatadze, G. I. Universal convergence of the specific volume changes of globular proteins upon unfolding. Biochemistry 2009, 48, 10846-10851. (51) Chalikian, T. V.; Breslauer, K. J. On volume changes accompanying conformational transitions of biopolymers. Biopolymers 1996, 39, 619-626. (52) Lin, L.-N.; Brandts, J. F.; Brandts, J. M.; Plotnikov, V. Determination of the volumetric properties of proteins and other solutes using pressure perturbation calorimetry. Anal. Biochem. 2002, 302, 144-160.

14 ACS Paragon Plus Environment

Page 15 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(53) Ravindra, R.; Winter, R. Pressure perturbation calorimetic studies of the solvation properties and the thermal unfolding of proteins in solution. Zeitschrift für Physikalische Chemie 2003, 217, 1221-1244. (54) Mitra, L.; Smolin, N.; Ravindra, R.; Royer, C.; Winter, R. Pressure perturbation calorimetric studies of the solvation properties and the thermal unfolding of proteins in solution-experiments and theoretical interpretation. Phys. Chem. Chem. Phys. 2006, 8, 1249-1265.

15 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 26

Figure Legends Figure 1. Pictorial representation of protein volume. The molecular surface (red line) encloses the solvent-excluded volume (VSE). The solvent-excluded volume (shaded area) consists of van der Waals volume (grey area, VvdW), and void volume (yellow area, VVoid). The solvent probes are shown as blue spheres. Figure 2. Temperature dependence of hydration coefficients kNP and kPol, indicating that kNP (black circles) has stronger temperature dependence than kPol (red inverted triangles). Open symbols show linearly extrapolated values. Error bars are standard error of the mean. Numerical values are given in Supplementary Table S1. Figure 3. Temperature dependence of native protein volume from 300 to 400K for eight different proteins. The increase in VTot (red inverse triangles) and VHyd (blue diamonds) with temperature is of similar magnitude, while VSE (black circles) only increases slightly due to the small increase in VVoid (green squares), indicating that VHyd makes a major contribution to the temperature dependence of VTot. Open symbols show linearly extrapolated values of hydration volume. Error bars are standard error of the mean and are of the size of the symbols. Figure 4. Dependence of the number of nearby atomic contacts on temperature in the native state of BPTI (black circles), Eglin C (blue squares), Ubiquitin (red triangles), Acylphosphatase (green inverted tringles), Ribonuclease (cyan diamonds), Lysozyme (pink hexagons), SNase (red squares), and SNase ∆PHS (open circles). Error bars are standard deviation. Figure 5. Characterization of volumetric properties of unfolded state ensemble of a 39-residue intrinsically-disordered peptide PAPf39. The increase in VTot (red inverse triangles) and VHyd (blue diamonds) with temperature is of similar magnitude, while VSE (black circles) only increases slightly due to the small increase in VVoid (green squares), indicating that VHyd makes a major contribution to the temperature dependence of VTot. Open symbols show linearly extrapolated values of hydration volume. Error bars are standard error of the mean and are of the size of the symbols.

Figure 6. Temperature dependence of unfolded protein volume across 300 to 400K range for eight different proteins. The increase in VTot (red inverse triangles) and VHyd (blue diamonds) with temperature is of similar magnitude, while VSE (black circles) only increases slightly due to the small increase in VVoid (green squares), indicating that VHyd makes a major contribution to the temperature dependence of VTot. Open symbols show linearly extrapolated values of hydration volume. Error bars are standard error of the mean and are of the size of the symbols. Figure 7. Temperature dependence of contributions from changes in void ∆VVoid (green squares) and hydration ∆VHyd (blue diamonds) volumes to the total volume change upon unfolding ∆VTot across the temperature range from 300 to 400K for 8 different proteins. ∆VVoid does not change significantly with temperature, indicating that ∆VHyd is a major contributor to the temperature dependence of the ∆VTot. Open symbols show linearly extrapolated values of hydration volume. Error bars are standard error of the mean. 16 ACS Paragon Plus Environment

Page 17 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 8. Comparison between calculated (black) and experimental (red) ∆ε, showing that the model can reasonably predict the relative differences of ∆ε between different proteins. Error bars are standard error of the mean.

17 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 1.

18 ACS Paragon Plus Environment

Page 18 of 26

Page 19 of 26

Figure 2.

0.6

KNP or Kpol (Å)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

0.5 0.4 0.3 0.2 0.1 0.0 280

320

360

400

Temperature (K)

19 ACS Paragon Plus Environment

The Journal of Physical Chemistry

300

380

3

340

380

300

Temperature (K)

380

300

380

380

SNase ∆PHS

21 20 19 18

VSE,N or VTot,N (Å3) x103

VSE,N or VTot,N (Å3) x103

340

VHyd,N or VVoid,N (Å3) x103

4 3 2 1

Temperature (K)

340

Temperature (K)

SNase

300

5 4 3 2 1

Lyz

20 19 18 17 16

Temperature (K)

22 21 20 19 18

380

VSE,N or VTot,N (Å3) x103

VSE,N or VTot,N (Å3) x103

340

VHyd,N or VVoid,N (Å3) x103

4 3 2 1 300

340

Temperature (K)

RNase

19 18 17 16 15

4 3 2 1

Acp

16 15 14 13 12

VSE,N or VTot,N (Å3) x103

3

VSE,N or VTot,N (Å ) x10

3

4 3 2 1

VHyd,N or VVoid,N (Å ) x10

3

Ubq

300

4 3 2 1

380

Temperature (K)

Temperature (K)

13 12 11 10 9

340

VHyd,N or VVoid,N (Å3) x103

340

VHyd,N or VVoid,N (Å3) x103

300

4 3 2 1

VHyd,N or VVoid,N (Å3) x103

4 3 2 1

Egl

12 11 10 9 8

VSE,N or VTot,N (Å3) x103

BPTI

10 9 8 7 6

VHyd,N or VVoid,N (Å3) x103

VSE,N or VTot,N (Å3) x103

Figure 3.

VHyd,N or VVoid,N (Å3) x103

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

300

340

380

Temperature (K)

20 ACS Paragon Plus Environment

Page 20 of 26

Page 21 of 26

Figure 4.

5000

Number of Contacts

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

4000

3000

2000

1000 300

320

340

360

380

400

Temperature (K)

21 ACS Paragon Plus Environment

The Journal of Physical Chemistry

Figure 5.

VSE,N or VTot,N (Å3) x103

PAPf39 7 6

2 1 0 300

340

VHyd,N or VVoid,N (Å3) x103

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

380

Temperature (K)

22 ACS Paragon Plus Environment

Page 22 of 26

Page 23 of 26

Figure 6.

3

3 3

VSE,U or VTot,U (Å ) x10

4 3 2 1

3

VSE,U or VTot,U (Å ) x10

Egl

12 11 10 9 8

VHyd,U or VVoid,U (Å ) x10

3 3

3

4 3 2 1

3

BPTI

10 9 8 7 6

VHyd,U or VVoid,U (Å ) x10

300

340

300

380

Ubq

340

380

Lyz

20 19 18 17 16

4 3 2 1

3

3

4 3 2 1

3

RNase

3

VSE,U or VTot,U (Å ) x10

3

380

VHyd,U or VVoid,U (Å ) x10

3

340

Temperature (K)

3

3

VHyd,U or VVoid,U (Å ) x10

VSE,U or VTot,U (Å ) x10

4 3 2 1 300

Temperature (K)

19 18 17 16 15

Acp

16 15 14 13 12

VSE,U or VTot,U (Å3) x103

VSE,U or VTot,U (Å3) x103

4 3 2 1 300

380

VHyd,U or VVoid,U (Å3) x103

13 12 11 10 9

340

Temperature (K)

Temperature (K)

VHyd,U or VVoid,U (Å3) x103

300

340

300

380

VSE,U or VTot,U (Å3) x103

340

380

Temperature (K)

SNase ∆PHS

22 21 20 19 18

4 3 2 1 300

340

VHyd,U or VVoid,U (Å3) x103

4 3 2 1 300

380

VSE,U or VTot,U (Å3) x103

SNase

22 21 20 19 18

340

Temperature (K)

Temperature (K)

VHyd,U or VVoid,U (Å3) x103

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

380

Temperature (K)

23 ACS Paragon Plus Environment

The Journal of Physical Chemistry

Figure 7.

0.4

0.5

0.2

0.0

0.0

-0.5

-0.2

-1.0

-0.4 300

340

3 3 ∆ VHyd or ∆ VVoid (Å ) x10

Egl 1.0

0.4

0.5

0.2

0.0

0.0

-0.5

-0.2

-1.0

-0.4

380

300

Temperature (K)

0

0.0 -0.2

-1

-0.4 -2

0.4 1

0.2

0

0.0 -0.2

-1

-0.4 -2

300

340

380

300

Temperature (K)

340

0.2

0

0.0

-1

-0.2

-2

-0.4 340

3 3 ∆ VHyd or ∆ VVoid (Å ) x10

1

2

0.4

1

0.2

0

0.0

-1

-0.2

-2

-0.4

380

300

Temperature (K)

0.2

0

0.0

-1

-0.2

-2

-0.4 380

Temperature (K)

2

0.4

1

0.2

0

0.0

-1

-0.2

-2

-0.4 300

340

380

Temperature (K)

24 ACS Paragon Plus Environment

3 3 ∆VTot (Å ) x10

1

3 3 ∆VTot (Å ) x10

0.4

3 3 ∆ VHyd or ∆ VVoid (Å ) x10

SNase ∆PHS

2

340

380

Temperature (K)

SNase

300

340

3 3 ∆VTot (Å ) x10

0.4 3 3 ∆VTot (Å ) x10

3 3 ∆ VHyd or ∆ VVoid (Å ) x10

Lyz

2

300

380

Temperature (K)

RNase

3 3 ∆ VHyd or ∆ VVoid (Å ) x10

3 3 ∆VTot (Å ) x10

0.2

3 3 ∆ VHyd or ∆ VVoid (Å ) x10

0.4 3 3 ∆VTot (Å ) x10

3 3 ∆ VHyd or ∆ VVoid (Å ) x10

Acp

2

1

380

Temperature (K)

Ubq

2

340

3 3 ∆VTot (Å ) x10

3 3 ∆ VHyd or ∆ VVoid (Å ) x10

BPTI 1.0

3 3 ∆VTot (Å ) x10

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 26

Page 25 of 26

Figure 8.

3

-1

∆ε (Å deg )

2

3

1

SN ase

Ly z

RN ase

Ac p

Ub q

0 Eg l

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

.

25 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

TOC Graphic

26 ACS Paragon Plus Environment

Page 26 of 26