Charged Surface Regulates the Molecular Interactions of

Mar 22, 2018 - LK has also a net charge of +1 contributed by 1 positive charge at the N-terminal. ... (39,50) When θ = 0, the dipole is perpendicular...
1 downloads 0 Views 3MB Size
Subscriber access provided by UNIV OF DURHAM

Biological and Environmental Phenomena at the Interface

Charged Surface Regulates the Molecular Interactions of Electrostatically Repulsive Peptides by Inducing Oriented Alignment Lin Zhang, and Yan Sun Langmuir, Just Accepted Manuscript • DOI: 10.1021/acs.langmuir.7b04308 • Publication Date (Web): 22 Mar 2018 Downloaded from http://pubs.acs.org on March 23, 2018

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 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 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.

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 31 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

Langmuir

1

Charged Surface Regulates the Molecular

2

Interactions of Electrostatically Repulsive Peptides

3

by Inducing Oriented Alignment

4

Lin Zhang, Yan Sun*

5

Department of Biochemical Engineering and Key Laboratory of Systems Bioengineering of the

6

Ministry of Education, School of Chemical Engineering and Technology, Tianjin University,

7

Tianjin 300072, China

8 9 10

ACS Paragon Plus Environment

1

Langmuir 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

1

ABSTRACT: Regulation of molecular orientation of charged dipeptides and involved

2

interactions by electrostatic repulsion from like-charged surfaces were studied using all-atom

3

molecular dynamics simulations. It was found that a charged surface can induce oriented

4

alignment of like-charged peptides, and the oriented alignment leads to enhanced electrostatic

5

repulsion between the peptide molecules. The findings are consistent with previous experimental

6

results about the inhibition of charged protein aggregation using like-charged ion-exchange resin.

7

Furthermore, the simulations provided molecular insights into this process, and demonstrated the

8

distinct regulation effect of like-charged surfaces on the molecular interactions between peptides

9

that possess an electric dipole structure. Both the charged surface and the electric dipole structure

10

of peptides were confirmed crucial for the regulation. The research are expected to facilitate the

11

rational design of surfaces or devices to regulate the behavior of amphoteric molecules such as

12

proteins for both in vivo and in vitro applications, which would contribute to the regulation of

13

protein-protein interactions and its application in life science and biotechnology.

Page 2 of 31

14 15

KEYWORDS: charged surface; like-charged peptide, oriented alignment; electrostatic

16

repulsion; molecular dynamics simulation

17

ACS Paragon Plus Environment

2

Page 3 of 31 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

1

Langmuir

1.

INTRODUCTION

2

Regulation of protein-protein interactions, protein adsorption and protein desorption

3

(repulsion) at various surfaces is of fundamental importance, because proteins are the most

4

abundant organic molecules in the cell and their interactions at surfaces are crucial in life science

5

and biotechnology,1-3 including but not limited to the biotransformation, transportation of matter,

6

energy transduction, cell-to-cell interactions and metabolic control.4,5 Molecular behavior of

7

proteins adsorbed at surfaces has been extensively studied because of its wide applications

8

ranging from material science to chemical biology,6-9 focusing on the orientation and

9

conformation of adsorbed proteins and consequent functionality.10-16 For instance, protein

10

adsorption at charged surfaces has been extensively investigated using molecular simulation,17-21

11

including the adsorption of charged protein to a like-charged surface.

12

Besides these adsorbed proteins, proteins repulsed from various surfaces, are also important in

13

both life sciences and bioengineering,22-26 but are not well understood. For instance, crowding

14

effect22-24 has long been considered effective in regulating protein interactions in vivo to assist

15

protein folding. The repulsive interactions between a protein and like-charged crowding agents

16

was speculated as a reason for the increased thermodynamic stability of the protein.23,27

17

However, possible enhancement on protein-protein interactions under crowded conditions is still

18

in dispute and thus an obstacle for the utilization of macromolecular crowding. Investigating the

19

influence of various surfaces on protein-protein interactions would be helpful but is usually

20

overlooked. Electrostatic forces in solutions of like-charged colloidal particles were investigated,

21

and electrostatic attraction was usually more focused.28-31 For electrostatic repulsion, a successful

22

regulation on protein-protein interactions using like-charged surfaces in vitro was experimentally

23

found32 in our previous work, which facilitated the on-pathway folding of proteins expressed as

ACS Paragon Plus Environment

3

Langmuir 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

1

inclusion bodies.33-35 The working mechanism was supposed as that the charged surfaces

2

oriented alignment of like-charged protein molecules inhibited protein-protein interactions.

3

However, experimental verification of such a mechanism, especially the electrostatic repulsion

4

oriented alignment36 was a challenge. Molecular dynamics (MD) simulation, as a powerful tool

5

that can offer clear microscopic information in a direct manner,6,37 even the orientational

6

alignment of proteins,38 was then used to explore the molecular insights into this process in our

7

previous work.39 Standing orientation of lysozyme molecules near a like-charged surface was

8

confirmed. However, complicated structure of protein molecules7,14,40,41 limited the investigation

9

on the oriented alignment at surfaces and consequent regulation on the molecular interactions.

10

Therefore, dipeptides were designed and used instead of protein molecules in this work.

11

Chemically, protein molecule is composed of one or more polypeptide chains with complicated

12

structures. The dipeptide has much simpler structure than that of a protein, but it can mimic the

13

amphoteric feature of proteins, which is expected important for the formation of orientation

14

controlled by electrostatic interactions. So the dipeptide was used as a dipole model of protein to

15

examine the protein orientation and molecular interactions at like-charged surface. Herein, four

16

dipeptides were designed, including KL, LK, LL, and LE, as shown in Figure 1a. Among these

17

peptides, both KL and LK are positively charged, but the former has larger electric dipole than

18

the latter; LL has several charged sites but is overall neutral; LE is negatively charged. Three

19

surfaces were designed for the simulations (Figure 1b), including one neutral, one positively

20

charged and one negatively charged. All-atom MD simulations were then performed to examine

21

the behavior of the peptides near the surfaces, focusing on the molecular orientations and

22

molecular interactions between peptides.

ACS Paragon Plus Environment

Page 4 of 31

4

Page 5 of 31 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

Langmuir

1 2

Figure 1. All-atom models of dipeptides (a) and surfaces (b). Four dipeptides were considered.

3

KL has a net charge of +1 at pH 7 because it has 2 positive charges and 1 negative charge. LK

4

has also a net charge of +1 contributed by 1 positive charge at the N-terminal. LL is neutral. LE

5

has a net charge of -1. Because the dipeptide has heterogeneous charge distribution, the atoms in

6

each dipeptide are divided into two groups, one includes all atoms with positive charge and the

7

other includes all atoms with negative charge. The center of mass of each group was calculated.

8

The equilibrated distance between these two groups was calculated to evaluate the electric

9

dipole, as shown below each peptide. All atoms were colored as CPK model and prepared using

10

Rasmol.42

11

2. MODELS AND METHODS

12

2.1. Model Construction. The AA models of dipeptides were constructed and the three

13

dimensional structures were determined using CharMM program (http://www.charmm-gui.org/),

14

as shown in Figure 1a. The net charge and electric dipole of each dipeptide are shown in Figure

15

1a.

16

To construct the simulation system in aqueous solution, two dipeptides were randomly placed

17

into a simulation box with a size of 6.0 × 6.0 × 6.0 nm3. The solvent molecules were then added.

ACS Paragon Plus Environment

5

Langmuir 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 31

1

The water molecule was treated using SPC model. The system was neutralized by adding Na+ or

2

Cl- as counter ions.

3

Charged surface was modeled as a neutral planar surface with immobilized positively or

4

negatively charged ligands, following the structure of ion-exchange resin.32 The surface is

5

composed of repetitive hexagonal unit43,44 with a side length of 0.152 nm (Figure 1b). For the

6

ligand, its topology was generated by the Dundee PRODRG 2.5 server (beta)45

7

(http://davapc1.bioch.dundee.ac.uk/cig-bin/prodrg_beta).

8

distribution were defined based on the parameters of similar structure in literature46 with minor

9

modification. Thereafter, ligands were immobilized to the matrix through a C-O-C bridge

10

(Figure 1b) with a ligand density of 0.21 mmol/mL according to the experimental data.32 In

11

positively/negatively charged surface, each ligand has a protonation state with a net charge of

12

+1/-1. Meanwhile, a surface with neutral ligands (with a net charge of zero) was constructed.

Its

charge

group

and

charge

13

To construct the simulation system containing charged/neutral surface, the surface was placed

14

at the bottom of a simulation box with a size of 7.3 × 6.3 × 6.0 nm3, and an inert surface was

15

placed at the top of the box to keep all molecules in the box. Two dipeptides equilibrated in

16

aqueous solution were used and put into the simulation box with the equilibrated conformations

17

in aqueous solution. That is, the dipeptides were firstly equilibrated in aqueous solution, and then

18

the equilibrated structures were used as the initial conformations at surface. Herein, rotation was

19

performed to ensure that each dipeptide had a same initial distance of 1.5 nm from the surface.

20

Solvent molecules around dipeptides in the equilibration stage were kept except those out of the

21

new simulation box. Supplement of solvent molecules was then performed to make correct

22

density in the simulation box. The system was neutralized by adding Na+ or Cl- as counter ions.

ACS Paragon Plus Environment

6

Page 7 of 31 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

Langmuir

1

These were then placed in the center of a cuboid box of size 7.3 × 6.3 × 100 nm3 for MD

2

simulation.

3

2.2. Simulation Method. MD simulations in the NVT ensemble were performed using

4

Gromacs 4.5.3 package (http://www.gromacs.org/)47,48 and Gromos96 43a1 force field.49 A 2 fs

5

time step was used to integrate the equations of motion. The cut-off of the Lennard-Jones (LJ)

6

potential was set to 1.2 nm. Standard PME function was used to deal with the Coulomb potential.

7

Periodic boundary was used in the x, y and z directions. Neighbor list was updated every 10

8

steps. The temperature was controlled at 298.15 K by the Berendsen method with a time constant

9

of 0.5 ps. The initial velocities of beads were generated according to the Maxwell distribution at

10

system temperature. Simulations were run for 20 ns in aqueous solution firstly, to examine the

11

behavior of dipeptides in bulk solution, and to obtain the equilibrated structures to provide the

12

initial conformations for the following 5 ns simulation at surfaces, as described in Section 2.1.

13

Sixty-four independent simulations were performed for each set of conditions. The snapshots

14

were prepared using Rasmol42 (http://www.umass.edu/microbio/rasmol/).

15

2.3. Dipeptide Orientation Analysis. The angle (θ) between the direction of dipeptide

16

dipole and z axis was calculated to evaluate the orientation of dipeptide, as reported in previous

17

work.39,50 When θ=0, the dipole is perpendicular to the surface with the negative pole facing the

18

surface. When θ=-90 or 90, the dipole is parallel to the surface (the dipole with positive x

19

coordinate is defined as plus, the other minus). θ=180° or -180° indicates a dipole also

20

perpendicular to the surface but with the positive pole facing the surface. A parameter P⊥ was

21

defined to represent the probability of dipeptide molecule with a θ between -30° to 30° to

22

evaluate the dipeptides perpendicular to the positively charged surface. It should be noted that P⊥

23

has been normalized by the volume fraction at certain θ in three dimensional space (see

ACS Paragon Plus Environment

7

Langmuir 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 8 of 31

1

Appendix S1, Supporting Information). Over negatively charged surface, P⊥ was defined to θ150° instead.

3

2.4. Dipeptide Movement Analysis. The z coordinate of the center of mass (com) of the

4

dipeptide (zcom) was monitored to depict the location of dipeptide. The distribution probability of

5

dipeptide, P(zcom,θ), was then calculated and shown by the filling colors. Herein, P(zcom,θ) was

6

calculated based on a statistics in the whole simulation to take all possible angle values. That is,

7

the dipeptide molecules with a certain zcom and θ were counted, and the obtained number was

8

divided by the total number of dipeptides sampled in the whole simulation to calculate P(zcom,θ).

9

Therefore, in the distribution map, at a certain zcom, the range of all possible θ was marked. The

10

conformational change of the peptides at the exclusion was evaluated using root mean square

11

deviation (RMSD) and radius of gyration (Rg). RMSD representing the structural change of a

12

molecule at time t with respect to a reference structure (herein the initial structure is used), was

13

calculated using g_rms provided by Gromacs. Rg representing the compactness of a molecule

14

was calculated using g_gyrate provided by Gromacs.

15

2.5. Molecular Interaction Analysis. The potential energies, including LJ potential and

16

Coulomb potential, were calculated using auxiliary program g_energy provided by Gromacs to

17

evaluate the molecular interactions between the dipeptide molecules or between the dipeptide

18

and the surface. Herein, hydrophobic interaction was considered and included in the LJ potential

19

energy. Hydrogen bond was analyzed using auxiliary program g_hbond provided by Gromacs.

20

The LJ potential between atoms i and j is calculated by Eq. (1).

VLJ (rij ) =

Cij12 rij12



Cij6

(1)

rij6

ACS Paragon Plus Environment

8

Page 9 of 31 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

Langmuir

rij

1

where

2

atom types.

3

is the distance between atoms i and j;

and

Cij6

are parameters depend on pairs of

The Coulomb potential between atoms i and j is calculated by Eq. (2). VC (rij ) = f

4

Cij12

qi q j

(2)

ε r rij

where f is electric conversion factor; ε r is dielectric constant; qi is the charge carried in atom i.

5 6

3. RESULTS AND DISCUSSION

7

3.1. Oriented Alignment of Dipeptides over Like-charged Surface. Two dipeptides,

8

equilibrated in aqueous solution, were put onto different surfaces with an initial distance of 1.5

9

nm to examine the regulation of protein behaviors by surfaces.

10

Quantitative description based on a statistics of 64 independent simulations was obtained

11

through P ⊥ and zcom values (Section 2.3). Low P ⊥ was observed in the aqueous solution,

12

indicating no preferred orientation but only random distribution (Figure S1a in Supporting

13

Information). At the neutral surface (Figure 2a), little change of zcom was observed, indicating

14

that the peptides randomly moved around their initial positions. The P⊥ values were small with

15

random changes, confirming that no oriented alignment was formed. This was confirmed by the

16

orientation and location of the peptides in typical snapshots (Figure S2) in a direct manner. No

17

obvious movement of the peptides was observed at a neutral surface (Figure S2a); KL moved

18

around the initial position and rotated randomly, and no preferred orientation was observed in the

19

snapshots. In contrary, at the like-charged surface, the P⊥ of KL changed from initial values to

20

approximately 1.0 immediately (within 0.02 ns), confirming the formation of oriented alignment

21

where the peptides were perpendicular to the surface with the negative pole facing the surface,

ACS Paragon Plus Environment

9

Langmuir 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

1

consistent with the snapshots in Figure S2b (snapshot at 0.02 ns). Slight conformational change

2

of the dipeptide was observed, as indicated by the increase of RMSD and Rg values (Figure S3).

3

It was attributed to the simultaneous electrostatic attraction at the negative pole of the dipeptide

4

and the electrostatic repulsion at the positive pole in the opposite directions, as a result of the

5

electric dipole structure. Meanwhile, the peptides were repulsed and then went far away, which

6

was consistent with the theoretical calculation of the equilibrated position of electric dipole near

7

a charged dot (see Appendix S2). In this process (rapid increase of zcom value), the P⊥ decreased

8

and reached an average value of 0.26 at 5 ns with fluctuations, indicating the diminishing of

9

oriented alignment following the fast exclusion of the peptides by the like-charged surface.

10

Decrease of Rg was also observed (Figure S3), indicating the diminishing of conformational

11

change, which is consistent with the protein behavior at desorption stage in chromatography in

12

our previous work.13,51-54

13

The distribution probability of the peptides was drawn for further evaluation of the orientation

14

based on a statistics of 64 independent simulations. Random distribution in aqueous solution was

15

observed (Figure S1b), indicated by the wide distribution with low probabilities. Similarly, wide

16

distribution of θ with low probabilities was observed at the neutral surface (Figure 2b),

17

indicating no preferred θ for the peptides. Meanwhile, the distribution region of the peptides at

18

the neutral surface is located at zcom = [0.5,1.5], which is near the initial value of zcom. That is, the

19

peptides maintained its initial state without significant change of orientation, consistent with the

20

snapshots in Figure S2. However, at the like-charged surface (Figure 2c), a dominant state with

21

zcom = [1.5,2.5] and θ = [-20°,20°] was observed, indicating the formation of oriented alignment

22

of the peptides. The preferred angle distribution near the like-charged surface located at the

23

range corresponding to the oriented perpendicular direction. As compared to the distribution

ACS Paragon Plus Environment

Page 10 of 31

10

Page 11 of 31 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

Langmuir

1

probability at neutral surface, an obvious change of peptides to approximately perpendicular

2

orientation can be concluded. However, this oriented alignment diminished with increased

3

distance from the surface. That is, the convergent angle distribution at small zcom became wider at

4

larger zcom (see the distribution in red, Figure 2c). According to the Coulomb potential [Eq. (2)],

5

the electrostatic interaction decreases with the increase of distance. Then the diminishing of

6

oriented alignment further highlighted the important role of electrostatic interaction from the

7

charged surface in this process.

8 9

Figure 2. Molecular behaviors of KL (a-c) and LE (d-f) at a neutral surface and a like-charged

10

surface. Time courses of P⊥ and zcom are shown in (a) for KL and (d) for LE, the distribution

11

probability near neutral surface represented as a zcom vs θ plot is shown in (b) for KL and (e) for

ACS Paragon Plus Environment

11

Langmuir 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

1

LE, and the distribution probability at the like-charged surface is shown in (c) for KL and (f) for

2

LE. The distribution probability is shown by the filling colors. Herein, because of the negatively

3

charged surface, the P⊥ for LE was defined as the probability of the dipeptide with a θ ≤ -150° or

4

≥ 150°, corresponding to an oriented alignment where the dipole is perpendicular to the surface

5

but the positive pole facing the surface. Schematic diagrams composed of representative

6

snapshots are provided in the figure top, where the dipoles of the dipeptides, at a direction from

7

red (negative pole) to blue (positive pole), is shown for better examination of their orientations.

8

The dipole moment is enlarged for clear view.

9

For further verification, the molecular behavior of the negatively charged LE at a negatively

10

charged surface was examined, as shown in Figures 2d to 2f. The P⊥ of peptides changed to 1.0

11

immediately (Figure 2d), indicating the formation of oriented alignment where the peptides were

12

perpendicular to the surface with the positive pole facing the surface, similar with the oriented

13

alignment of KL over the positively charged surface (Figure 2a). Thereafter, P⊥ decreased with

14

the rapid increase of zcom values, and reached an average value of 0.18 at 5 ns with fluctuations,

15

indicating the diminishing of oriented alignment following the fast exclusion of peptides by the

16

like-charged surface. In the distribution probability of LE at negatively charged surface (Figure

17

2f), a dominant state with zcom = [1.7,2.5] and θ ≤ -170° or ≥ 160° was observed, proving the

18

formation of like-charged surface oriented alignment.

19

3.2. Regulation of Intermolecular Interactions by Charged Surface. Molecular

20

interactions between dipeptides were monitored, including the electrostatic interactions,

21

hydrophobic interactions and hydrogen bonds. However, the value of LJ potential energy was

22

much smaller than that of Coulomb potential energy (see Figure S4), because the LJ potential

23

energy decreases much faster with the increase of distance, as shown in Eqs. (1) and (2).

ACS Paragon Plus Environment

Page 12 of 31

12

Page 13 of 31 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

Langmuir

1

Hydrogen bond is also only effective within a small distance of about 0.35 nm. Therefore,

2

hydrophobic interactions and hydrogen bonds were not discussed herein. The Coulomb potential

3

energy and the distance between two dipeptides were used to evaluate the effect of charged

4

surface on the molecular interactions between the two peptides. Herein, △ d=d-d0 was used

5

instead of d to avoid the disturbance of initial distance between dipeptides (d0). EC around zero

6

was observed with significant fluctuations in aqueous solution (Figure S1c), indicating random

7

collision of the peptides in bulk solution. At the neutral surface (Figure 3), EC had significant

8

fluctuations near zero but without a definite trend, indicating that random movement did not

9

influence the molecular interactions between the peptides. In contrary, at the like-charged surface,

10

for both positively charged KL (Figure 3a) and negatively charged LE (Figure 3b), immediate

11

increases of EC were observed, indicating that the oriented alignments (Figure 2) indeed

12

enhanced the repulsion between the peptides, which was also confirmed by the larger △d than

13

that at the neutral surface. However, as the discussion refer to Figure 1, the oriented alignment

14

diminished when the dipeptides went away from the surface, because the electrostatic interaction

15

decreased with increased distance according to the Coulomb potential [Eq. (2)]. As a result, the

16

enhanced repulsion between the peptides diminished. △ d of KL (LE) dipeptides near like-

17

charged surface decreased and became similar as that above neutral surface at the last part of

18

simulation. This further confirmed that the like-charged surface oriented alignment was the

19

reason for increased molecular repulsion between the peptides, and this can explain the

20

experimental observation of reduced protein aggregation in a suspension with like-charged

21

particles.32,35,55,56 From the experimental results, enhanced refolding of charged protein at like-

22

charged surface was confirmed. However, the exploration of molecular mechanism was still a

23

challenge using experimental approaches. From the simulation results, the like-charged surface

ACS Paragon Plus Environment

13

Langmuir 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

1

oriented alignment and consequent enhancement of electrostatic repulsion between the

2

dipeptides were confirmed. That is, the simulation results can explore the molecular mechanism

3

about the function of charged surface directly. Based on these findings, charged surface can be

4

designed and used to induce microscopic oriented alignment of peptides, which leads to the

5

enhanced electrostatic repulsion between peptides, and accounted for the macroscopic

6

experimental results of reduced protein aggregation by like-charged particles.

Page 14 of 31

7 8

Figure 3. Molecular interactions between KL (a) or LE (b) at a neutral or a like-charged surface,

9

as shown by the time courses of EC and △d. △d is the increase of d as compared to the initial

10

value. Schematic diagrams composed of representative snapshots are provided in the figure top.

11

3.3. The Dependence of Oriented Alignment on Electric Dipole. As mentioned above, a

12

like-charged surface was necessary to realize the regulation on protein orientation. Besides this,

13

the electric dipole of peptides may also be crucial. Simulations using other dipeptides with

14

different dipoles and net charges were therefore performed to evaluate the contribution of the

ACS Paragon Plus Environment

14

Page 15 of 31 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

Langmuir

1

electric dipole and net charge. LK with the same net charge of +1 as KL but lower charge density

2

was used, together with a neutral dipeptide LL (see Figure 1).

3 4

Figure 4. Molecular behaviors of LK and LL over a positively charged surface, where the time

5

courses of P⊥ and zcom are shown in (a), the distribution probability represented as zcom vs θ plot

6

is shown in (b) for LK and (c) for LL, and the time courses of EC and △d values are shown in

7

(d).

8

The P⊥ and zcom were calculated, as shown in Figure 4a. Similar results were observed for LK

9

as those for KL (Figure 2). Oriented alignment was formed immediately and then diminished in

10

the exclusion process of LK. For LL, however, a significant difference was observed. The

ACS Paragon Plus Environment

15

Langmuir 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

1

peptides were adsorbed onto the surface very quickly rather than being repulsed, as indicated by

2

the zcom. This was consistent with the theoretical calculation in Appendix S2 (Supporting

3

Information). Obvious oriented alignment of the adsorbed LL molecules was observed at the

4

surface. That was attributed to the heterogeneous charge distribution of LL. That is, LL should

5

also be considered as an electric dipole with distinct negatively charged part (negative pole) and

6

positively charged part (positive pole). The attraction of negative part at surface induced the

7

formation of oriented alignment. Moreover, LL has an overall neutral charge, so the attraction

8

became stronger due to smaller distance as compared with the repulsion of positively charged

9

part (positive pole), leading to the adsorption rather than repulsion of the whole LL molecule.

10

The distribution probability was calculated to evaluate the orientation of the peptides at the

11

charged surface based on a statistics of 64 independent simulations, as shown by the filling

12

colors in Figures 4b and 4c. For LK, a dominant distribution located in the region of θ = [-30°,

13

30°] was observed, indicating the formation of a perfect oriented alignment where the peptides

14

were perpendicular to the surface with the negative pole facing the surface. More converged state

15

around θ = 0° was observed for LL with a small zcom, indicating the formation of oriented

16

alignment on the surface as well as the dominant attraction. The more converged state further

17

confirmed the enhanced electrostatic attraction due to the small distance from the surface.

18

The Coulomb potential energy and the distance between two dipeptides were monitored to

19

examine the contribution of charged surface on inhibiting the aggregation of the peptides, as

20

shown in Figure 4d. For LK, rapid increase of △d was observed, followed by an immediate

21

increase of EC, indicating the oriented alignment indeed enhanced the repulsion between the

22

peptides. As compared to KL (Figure 3a), however, the increase of EC was smaller. This clearly

23

indicates that the smaller dipole resulted in a weaker enhancement of repulsion between the

ACS Paragon Plus Environment

Page 16 of 31

16

Page 17 of 31 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

Langmuir

1

peptides at like-charged surface. For LL, the increase of EC was also observed, but without

2

significant increase of △d, because the peptides were adsorbed onto the surface and could not

3

move freely (see zcom in Figure 4a). Similar results were obtained for the positively charged KL

4

at a negatively charged surface (see Figure S5).

5

Therefore, the results indicated that the net charge of protein determines the adsorption or

6

repulsion at a charged surface. Charged protein is repulsed from a like-charged surface, such as

7

the repulsion of positively charged KL by a positively charged surface, or negatively charged LE

8

by a negatively charged surface (Figure 2). In contrary, protein without effective repulsion by the

9

surface will be adsorbed, such as the adsorption of neutral LL on a positively charged surface

10

(Figure 4), or positively charged KL at a negatively charged surface (Figure S5). The

11

dependence of adsorption or repulsion on the net charge of protein is consistent with the

12

theoretical calculation in Appendix S2 (Supporting Information). Moreover, the regulation of

13

intermolecular interactions of adsorbed peptides was different from that for repulsed peptides.

14

Because the electrostatic interaction decreased with increased distance according to the Coulomb

15

potential [Eq. (2)], the diminishing of regulation was observed during the excluding process of

16

repulsed peptides, as the discussion refer to Figure 1. For the adsorbed peptides, the maintained

17

small distance from the surface caused large electrostatic interaction and strong regulation. For

18

example, LL adsorbed at the surface make it obtain a large P ⊥ above 0.6 and a large △ d,

19

indicating a good spatial isolation of peptides. The spatial isolation of proteins on ion-exchange

20

chromatography (IEC) was widely described in experimental results, and was considered as a

21

driving force for the protein refolding in IEC.57-59 Protein adsorption driven by electrostatic

22

attraction from the IEC resin is the key to separate the bound protein molecules and helpful for

23

the protein refolding.

ACS Paragon Plus Environment

17

Langmuir 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

1

Moreover, the results indicated that the electric dipole of a protein is crucial for the formation

2

of charged surface oriented alignment. A protein with heterogeneous charge distribution can

3

possess preferred orientation at a charged surface. As evidenced by the results in the present

4

study, all the four dipeptides, KL, LK, LL and EK, can form oriented alignment at a charged

5

surface due to their electric dipoles, regardless of being adsorbed or repulsed. The attraction from

6

the surface to the part of a dipeptide with opposite charge is the determinant for the preferred

7

adsorption sites. For example, the negative pole of LL determined the preferred orientation of LL

8

and consequent oriented alignment. The adsorption of protein on a charged surface due to its

9

heterogeneous surface has been extensively reported in the simulation results.17-21 The preferred

10

adsorption sites and orientation have been discussed. The simulation results of LL on a charged

11

surface, featured by the preferred adsorption sites and oriented alignment, are consistent with the

12

results reported in literature.57-59 Moreover, the regulation of molecular interactions between

13

peptides by the oriented alignment was highlighted herein, which accounted for the experimental

14

observation about the inhibition of protein aggregation by like-charged resins.32 This may be the

15

reason for protein folding rather than aggregation in crowded conditions, which can interpret the

16

crowding effect. In crowded conditions, proteins are always near various surfaces (including

17

various charged surfaces) provided by the crowding agents. The induced oriented alignment of

18

protein by such charged surfaces can be expected due to the heterogeneous charge distribution

19

(electric dipole) of protein molecules, which would then enhance the repulsion between proteins

20

and thus help enhance the inhibition of their aggregation and improve the on-pathway protein

21

folding.

Page 18 of 31

22 23

ACS Paragon Plus Environment

18

Page 19 of 31 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

1

Langmuir

4. CONCLUSIONS

2

In summary, charged surfaces can induce the formation of oriented alignment of peptides,

3

leading to enhanced electrostatic repulsion between peptides or proteins, which is considered as

4

the reason for the inhibition of aggregation and the enhancement of on-pathway protein folding.

5

The oriented alignment of peptides diminished with increasing distance from the like-charged

6

surface, and can be enhanced by increasing electric dipole of the peptides. The simulation results

7

explained the experimental observation of reduced protein aggregation in a suspension with like-

8

charged particles, and confirmed the like-charged surface oriented alignment and consequent

9

enhanced molecular repulsion between the peptides at molecular level. These results have

10

explored the working mechanism of charged surfaces on the regulation of molecular interactions

11

between peptides, which would be helpful for interpreting the crowding effect in vivo and the

12

enhanced on-pathway folding of high concentration proteins by like-charged additives in vitro.

13

The research will thus facilitate the rational design of surfaces or devices for regulating the

14

interfacial behaviors of protein molecules both in vitro and in vivo.

15 16

ASSOCIATED CONTENT

17

Supporting Information.

18

Appendix about the derivation of volume fraction in three dimensional space, appendix about the

19

calculation of the equilibrated position of electric dipole near a charged dot, molecular behaviors

20

of KL and LE in aqueous solution, snapshots of two KL dipeptides over a neutral surface and a

21

charged surface, LJ potential energies of KL and LE dipeptides at a like-charged surface, the

22

time courses of RMSD and Rg of KL and LE dipeptides at a like-charged surface, and molecular

ACS Paragon Plus Environment

19

Langmuir 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

1

behaviors of KL at a negatively charged surface. This material is available free of charge via the

2

Internet at http://pubs.acs.org.

3

AUTHOR IMFORMATION

4

Corresponding author

5

*Tel & Fax: +86 22 27403389; E-mail addresses: [email protected] (Y. Sun).

6

Author Contributions

7

The manuscript was written through contributions of all authors. All authors have given approval

8

to the final version of the manuscript.

9

Notes

10

The authors declare no competing financial interest.

11

ACKNOWLEDGMENTS

12

This work was supported by the Natural Science Foundation of China (Nos. 91534119,

13

21236005, 21376173 and 21621004) and the Innovation Foundation of Tianjin University.

14 15

REFERENCES

16

(1) Belfort, G.; Lee, C. S. Attractive and Repulsive Interactions Between and within Adsorbed Ribonuclease a

17 18 19 20 21 22 23

Page 20 of 31

Layers. P. Natl. Acad. Sci. U.S.A. 1991, 88, 9146–9150. (2) Scott, D. E.; Bayly, A. R.; Abell, C.; Skidmore, J. Small Molecules, Big Targets: Drug Discovery Faces the Protein-Protein Interaction Challenge. Nat. Rev. Drug Discov. 2016, 15, 533–550. (3) Keskin, O.; Gursoy, A.; Ma, B.; Nussinov, R. Principles of Protein-Protein Interactions: What are the Preferred Ways for Proteins to Interact? Chem. Rev. 2008, 108, 1225–1244. (4) Yang, H.; Yuan, B.; Zhang, X.; Scherman, O. A. Supramolecular Chemistry at Interfaces: Host-Guest Interactions for Fabricating Multifunctional Biointerfaces. Accounts Chem. Res. 2014, 47, 2106–2115.

ACS Paragon Plus Environment

20

Page 21 of 31 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

1 2 3 4 5 6 7 8

Langmuir

(5) Rao, V. S.; Srinivas, K.; Sujini, G. N.; Kumar, G. N. S. Protein-Protein Interaction Detection: Methods and Analysis. International Journal of Proteomics 2014, 2014, 147648–147648. (6) Ozboyaci, M.; Kokh, D. B.; Corni, S.; Wade, R. C. Modeling and Simulation of Protein-Surface Interactions: Achievements and Challenges. Q. Rev. Biophys. 2016, 49, 1–45. (7) Zhang, L.; Sun, Y. Molecular Simulation of Adsorption and its Implications to Protein Chromatography: A Review. Biochem. Eng. J. 2010, 48, 408–415. (8) Duval, F.; van Beek, T. A.; Zuilhof, H. Key Steps Towards the Oriented Immobilization of Antibodies Using Boronic Acids. Analyst 2015, 140, 6467–6472.

9

(9) Murugan, P.; Krishnamurthy, M.; Jaisankar, S. N.; Samanta, D.; Mandal, A. B. Controlled Decoration of the

10

Surface with Macromolecules: Polymerization On a Self-Assembled Monolayer (Sam). Chem. Soc. Rev. 2015,

11

44, 3212–3243.

12

(10) Penna, M. J.; Mijajlovic, M.; Biggs, M. J. Molecular-Level Understanding of Protein Adsorption at the

13

Interface Between Water and a Strongly Interacting Uncharged Solid Surface. J. Am. Chem. Soc. 2014, 136,

14

5323–5331.

15 16

(11) Puddu, V.; Perry, C. C. Peptide Adsorption On Silica Nanoparticles: Evidence of Hydrophobic Interactions. ACS Nano 2012, 6, 6356–6363.

17

(12) Lin, D. Q.; Tong, H. F.; Wang, H. Y.; Shao, S.; Yao, S. J. Molecular Mechanism of Hydrophobic Charge-

18

Induction Chromatography: Interactions Between the Immobilized 4-Mercaptoethyl-Pyridine Ligand and Igg.

19

J. Chromatogr. A 2012, 1260, 143–153.

20

(13) Zhang, L.; Zhao, G. F.; Sun, Y. Molecular Insight Into Protein Conformational Transition in Hydrophobic

21

Charge Induction Chromatography: A Molecular Dynamics Simulation. J. Phys. Chem. B 2009, 113, 6873–

22

6880.

23 24 25 26 27 28

(14) Watkins, A. M.; Bonneau, R.; Arora, P. S. Side-Chain Conformational Preferences Govern Protein-Protein Interactions. J. Am. Chem. Soc. 2016, 138, 10386–10389. (15) Fears, K. P.; Clark, T. D.; Petrovykh, D. Y. Residue-Dependent Adsorption of Model Oligopeptides On Gold. J. Am. Chem. Soc. 2013, 135, 15040–15052. (16) McUmber, A. C.; Randolph, T. W.; Schwartz, D. K. Electrostatic Interactions Influence Protein Adsorption (but Not Desorption) at the Silica-Aqueous Interface. J. Phys. Chem. Lett. 2015, 6, 2583–2587.

ACS Paragon Plus Environment

21

Langmuir 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

1 2

(17) Liu, J.; Peng, C.; Yu, G.; Zhou, J. Molecular Simulation Study of Feruloyl Esterase Adsorption On Charged Surfaces: Effects of Surface Charge Density and Ionic Strength. Langmuir 2015, 31, 10751–10763.

3

(18) Peng, C.; Liu, J.; Xie, Y.; Zhou, J. Molecular Simulations of Cytochrome C Adsorption On Positively Charged

4

Surfaces: The Influence of Anion Type and Concentration. Phys. Chem. Chem. Phys. 2016, 18, 9979–9989.

5

(19) Romanowska, J.; Kokh, D. B.; Wade, R. C. When the Label Matters: Adsorption of Labeled and Unlabeled

6 7 8

Proteins On Charged Surfaces. Nano Lett. 2015, 15, 7508–7513. (20) Kubiak-Ossowska, K.; Mulheran, P. A. Mechanism of Hen Egg White Lysozyme Adsorption On a Charged Solid Surface. Langmuir 2010, 26, 15954–15965.

9

(21) Kubiak-Ossowska, K.; Jachimska, B.; Mulheran, P. A. How Negatively Charged Proteins Adsorb to Negatively

10

Charged Surfaces: A Molecular Dynamics Study of Bsa Adsorption On Silica. J. Phys. Chem. B 2016, 120,

11

10463–10468.

12 13 14 15 16 17 18 19

(22) Guigas, G.; Weiss, M. Effects of Protein Crowding On Membrane Systems. Bba.-Biomembranes 2016, 1858, 2441–2450. (23) Mittal, S.; Chowhan, R. K.; Singh, L. R. Macromolecular Crowding: Macromolecules Friend Or Foe. Bba.Gen. Subjects 2015, 1850, 1822-1831. (24) Sapir, L.; Harries, D. Is the Depletion Force Entropic? Molecular Crowding Beyond Steric Interactions. Curr. Opin. Colloid in 2015, 20, 3–10. (25) Shao, Q.; Jiang, S. Y. Molecular Understanding and Design of Zwitterionic Materials. Adv. Mater. 2015, 27, 15–26.

20

(26) Zheng, J.; Li, L. Y.; Tsao, H. K.; Sheng, Y. J.; Chen, S. F.; Jiang, S. Y. Strong Repulsive Forces Between

21

Protein and Oligo (Ethylene Glycol) Self-Assembled Monolayers: A Molecular Simulation Study. Biophys. J.

22

2005, 89, 158–166.

23 24 25 26 27 28

Page 22 of 31

(27) Sarkar, M.; Lu, J.; Pielak, G. J. Protein Crowder Charge and Protein Stability. Biochemistry-US 2014, 53, 1601–1606. (28) Linse, P.; Lobaskin, V. Electrostatic Attraction and Phase Separation in Solutions of Like-Charged Colloidal Particles. Phys. Rev. Lett. 1999, 83, 4208–4211. (29) Angelescu, D. G.; Linse, P. Monte Carlo Simulation of the Mean Force Between Two Like-Charged Macroions with Simple 1 : 3 Salt Added. Langmuir 2003, 19, 9661–9668.

ACS Paragon Plus Environment

22

Page 23 of 31 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

Langmuir

1

(30) Tiraferri, A.; Maroni, P.; Borkovec, M. Adsorption of Polyelectrolytes to Like-Charged Substrates Induced by

2

Multivalent Counterions as Exemplified by Poly(Styrene Sulfonate) and Silica. Phys. Chem. Chem. Phys.

3

2015, 17, 10348–10352.

4

(31) Luque-Caballero, G.; Martin-Molina, A.; Quesada-Perez, M. Polyelectrolyte Adsorption Onto Like-Charged

5

Surfaces Mediated by Trivalent Counterions: A Monte Carlo Simulation Study. J. Chem. Phys. 2014, 140,

6

174701.

7 8 9 10 11 12

(32) Wang, G.; Dong, X.; Sun, Y. Ion–Exchange Resins Greatly Facilitate Refolding of Like–Charged Proteins at High Concentrations. Biotechnol. Bioeng. 2011, 108, 1068–1077. (33) Yu, L. L.; Dong, X. Y.; Sun, Y. Ion-Exchange Resins Facilitate Like-Charged Protein Refolding: Effects of Porous Solid Phase Properties. J. Chromatogr. A 2012, 1225, 168–173. (34) Yang, C. Y.; Li, M.; Dong, X. Y.; Sun, Y. A Double-Modification Strategy for Enhancing Charge Density of Mono-Sized Beads for Facilitated Refolding of Like-Charged Protein. J. Chromatogr. A 2013, 1299, 85–93.

13

(35) Liu, H.; Dong, X.; Sun, Y. Grafting Iminodiacetic Acid On Silica Nanoparticles for Facilitated Refolding of

14

Like-Charged Protein and its Metal-Chelate Affinity Purification. J. Chromatogr. A 2016, 1429, 277–283.

15

(36) Liu, B.; Besseling, T. H.; Blaaderen, A. V.; Imhof, A. Confinement Induced Plastic Crystal-to-Crystal

16 17 18 19 20 21 22 23 24 25 26

Transitions in Rodlike Particles with Long-Ranged Repulsion. Phys. Rev. Lett. 2015, 115, 078301. (37) Akimov, A. V.; Prezhdo, O. V. Large-Scale Computations in Chemistry: A Bird's Eye View of a Vibrant Field. Chem. Rev. 2015, 115, 5797–5890. (38) Schröder, C.; Steinhauser, O.; Sasisanker, P.; Weingärtner, H. Orientational Alignment of Amyloidogenic Proteins in Pre-Aggregated Solutions. Phys. Rev. Lett. 2015, 114, 128101. (39) Bai, S.; Li, H.; Zhang, L. Standing Orientation of Lysozymes Induced by Electrostatically Repulsive Surface. Acta Phys.-Chim. Sin. 2013, 29, 849–857. (40) Zhang, L.; Lua, L.; Middelberg, A.; Sun, Y.; Connors, N. K. Biomolecular Engineering of Virus-Like Particles Aided by Computational Chemistry Methods. Chem. Soc. Rev. 2015, 44, 8608–8618. (41) Dijk, E. V.; Varilly, P.; Knowles, T. P.; Frenkel, D.; Abeln, S. Consistent Treatment of Hydrophobicity in Protein Lattice Models Accounts for Cold Denaturation. Phys. Rev. Lett. 2016, 116, 078101.

27

(42) Sayle, R.; Milnerwhite, E. Rasmol - Biomolecular Graphics for All. Trends Biochem. Sci. 1995, 20, 374–376.

28

(43) Dresselhaus, M. S.; Avouris, P. Introduction to Carbon Materials Research. Top. Appl. Phys. 2001, 80, 1–9.

ACS Paragon Plus Environment

23

Langmuir 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

1 2 3 4

(44) Hummer, G.; Rasaiah, J. C.; Noworyta, J. P. Water Conduction through the Hydrophobic Channel of a Carbon Nanotube. Nature 2001, 414, 188–190. (45) Schuttelkopf, A. W.; van, A. D. Prodrg: A Tool for High-Throughput Crystallography of Protein-Ligand Complexes. Acta Crystallographica Section D-Biological Crystallography 2004, 60, 1355–1363.

5

(46) Chandrasekhar, I.; Kastenholz, M.; Lins, R. D.; Oostenbrink, C.; Schuler, L. D.; Tieleman, D. P.; van

6

Gunsteren, W. F. A Consistent Potential Energy Parameter Set for Lipids: Dipalmitoylphosphatidylcholine as a

7

Benchmark of the Gromos96 45a3 Force Field. Eur. Biophys. J. Biophy. 2003, 32, 67–77.

8 9 10 11 12 13 14 15 16 17

(47) Berendsen, H. J.; Vanderspoel, D.; Vandrunen, R. Gromacs - A Message-Passing Parallel Molecular-Dynamics Implementation. Comput. Phys. Commun. 1995, 91, 43–56. (48) Lindahl, E.; Hess, B.; van, S. D. Gromacs 3.0: A Package for Molecular Simulation and Trajectory Analysis. J. Mol. Model. 2001, 7, 306–317. (49) Mackerell, A. D. Empirical Force Fields for Biological Macromolecules: Overview and Issues. J. Comput. Chem. 2004, 25, 1584–1604. (50) Yu, G.; Liu, J.; Zhou, J. Mesoscopic Coarse-Grained Simulations of Hydrophobic Charge Induction Chromatography (Hcic) for Protein Purification. AIChE J. 2015, 61, 2035–2047. (51) Zhang, L.; Zhao, G. F.; Sun, Y. Effects of Ligand Density On Hydrophobic Charge Induction Chromatography: Molecular Dynamics Simulation. J. Phys. Chem. B 2010, 114, 2203–2211.

18

(52) Zhang, L.; Bai, S.; Sun, Y. Molecular Dynamics Simulation of the Effect of Ligand Homogeneity On Protein

19

Behavior in Hydrophobic Charge Induction Chromatography. Journal of Molecular Graphics and Modelling

20

2010, 28, 863–869.

21

(53) Zhang, L.; Zhao, G. F.; Sun, Y. Molecular Dynamics Simulation and Experimental Validation of the Effect of

22

Ph On Protein Desorption in Hydrophobic Charge Induction Chromatography. Mol. Simulat. 2010, 36, 1096–

23

1103.

24

(54) Zhang, L.; Sun, Y. Effect of Ligand Chain Length On Hydrophobic Charge Induction Chromatography

25

Revealed by Molecular Dynamics Simulations. Frontiers of Chemical Science and Engineering 2013, 7, 456–

26

463.

27 28

Page 24 of 31

(55) Yang, C. Y.; Yu, L. L.; Dong, X. Y.; Sun, Y. Mono-Sized Microspheres Modified with Poly(Ethylenimine) Facilitate the Refolding of Like-Charged Lysozyme. React. Funct. Polym. 2012, 72, 889–896.

ACS Paragon Plus Environment

24

Page 25 of 31 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

1 2 3 4 5 6 7 8

Langmuir

(56) Yang, C.; Dong, X.; Sun, Y. Mechanistic Studies of Protein Refolding Facilitated by Like-Charged Polymers. React. Funct. Polym. 2013, 73, 1405–1411. (57) Langenhof, M.; Leong, S.; Pattenden, L. K.; Middelberg, A. Controlled Oxidative Protein Refolding Using an Ion-Exchange Column. J. Chromatogr. A 2005, 1069, 195–201. (58) Machold, C.; Schlegl, R.; Buchinger, W.; Jungbauer, A. Matrix Assisted Refolding of Proteins by Ion Exchange Chromatography. J. Biotechnol. 2005, 117, 83–97. (59) Freydell, E. J.; van der Wielen, L.; Eppink, M.; Ottens, M. Ion-Exchange Chromatographic Protein Refolding. J. Chromatogr. A 2010, 1217, 7265–7274.

9

ACS Paragon Plus Environment

25

Langmuir 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 26 of 31

Table of contents only

ACS Paragon Plus Environment

26

Page 27 of 31 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

Langmuir

Figure 1 64x23mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir 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 2 134x102mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 28 of 31

Page 29 of 31 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

Langmuir

Figure 3 94x49mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir 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 4 133x99mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 30 of 31

Page 31 of 31 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

Langmuir

TOC 56x37mm (300 x 300 DPI)

ACS Paragon Plus Environment