Structure and Relaxation in Solutions of ... - ACS Publications

E-mail: [email protected]. Phone: 617-324-7359. 1. Page 1 of 41. ACS Paragon Plus Environment. The Journal of Physical Chemistry. 1. 2. 3. 4. 5. 6. 7. 8. ...
0 downloads 10 Views 10MB Size
Subscriber access provided by MT ROYAL COLLEGE

Article

Structure and Relaxation in Solutions of Monoclonal Antibodies Gang Wang, Zsigmond Varga, Jennifer Hofmann, Isidro E Zarraga, and James W. Swan J. Phys. Chem. B, Just Accepted Manuscript • DOI: 10.1021/acs.jpcb.7b11053 • Publication Date (Web): 22 Feb 2018 Downloaded from http://pubs.acs.org on February 25, 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.

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 41 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

Structure and Relaxation in Solutions of Monoclonal Antibodies Gang Wang,





Zsigmond Varga,

Jennifer Hofmann,

James W. Swan

†Department





Isidro E. Zarraga,

and

∗,†

of Chemical Engineering, Massachusetts Institute of Technology, Cambridge, MA

‡Late

Stage Pharmaceutical Development, Genentech Inc., South San Francisco, CA

E-mail: [email protected]

Phone: 617-324-7359

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

Abstract Reversible self-association of therapeutic antibodies is a key factor in high protein solution viscosities. In the present work, a coarse-grained computational model accounting for electrostatic, dispersion and long-ranged hydrodynamic interactions of two model monoclonal antibodies is applied to understand the nature of self-association, predicting the solution micro-structure and resulting transport properties of the solution. For the proteins investigated, the structure factor across a range of solution conditions shows quantitative agreement with neutron scattering experiments. We observe homogeneous, dynamical association of the antibodies with no evidence of phase separation. Calculations of self-diusivity and viscosity from coarse-grained dynamic simulations show the appropriate trends with concentration, but respectively over and under predict the experimentally measured values. By adding constraints to the self-associated clusters that rigidify them under ow, prediction of the transport properties is signicantly improved with respect to experimental measurements. We hypothesize that these rigidity constraints are associated with missing degrees of freedom in the coarse-grained model resulting from patchy and heterogeneous interactions among coarse-grained domains. These results demonstrate how structural anisotropy and anisotropy of interactions generated by features at the 2-5 nm length scale in antibodies are sucient to recover the dynamics and rheological properties of these important macromolecular solutions.

Introduction Therapeutic protein products with high concentration (>50 mg/ml) have presented researchers with a crucial yet challenging problem in the biopharmaceutical industry. Therapeutic monoclonal antibodies (MAbs) featuring high anity and specicity for antigens, are widely used for immunotherapy such as novel cancer therapies. In general, high doses of the antibodies are required. Because the antibody formulations are not concentrated enough for subcutaneous injection, expensive and time-consuming intravenous administration is often used. 1 Subcutaneous administration has an upper limit of dosage volume, demanding 2

ACS Paragon Plus Environment

Page 2 of 41

Page 3 of 41 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

formulations with concentrations as high as 100 mg/ml. 2 However, MAb solutions with concentrations at that level, depending on the peptide sequence, often possess extremely high viscosities, 36 and thus lead to diculties in processing and delivery. Challenges also arise in terms of stability, such as irreversible protein aggregation, adding to the risk of lost ecacy and induced immunogenicity. 5,7 Predicting protein solution viscosity from its primary structure is an important scientic and engineering challenge waiting to be met. Over the past few years, a large body of work has investigated the solution properties of various monoclonal antibodies under dierent solution conditions, such as protein concentration, pH, and ionic strength. 3,6,8 It has been demonstrated that self-association among antibodies is a key factor to elevated solution viscosity. 3 Yadav et al. investigated the eect of charge distribution on self-association by utilizing rheology and dynamic light scattering (DLS) on two types of antibodies and their charge-swap mutants. 9 The two MAbs, namely MAb1 and MAb2, have 92% sequence similarity, but were observed to have dramatically dierent viscosities. 6 Electrostatic interactions were assumed to be the determining factor of self-association, since adding NaCl eectively decreased the solution viscosity. This was further conrmed by performing all-atom molecular dynamics (MD) simulation on a single molecule and Poisson-Boltzmann (PB) calculations to understand the molecular level electrostatics. 10 A recent study applied small-angle neutron scattering (SANS) to probe the solution structure and protein-protein interactions (PPI). MAb1 was demonstrated to exhibit specic anisotropic interactions between the molecules. 11 Short-time diusivities of MAb1 and MAb2 were measured by neutron spin echo (NSE) experiments and the dierence was attributed to the formation of self-associated clusters in MAb1 solutions. However, neutron scattering only serves as an indirect method to probe the structure of antibody solutions. It presents the inverse problem of interpreting data and does not provide direct access to the solution structure. It remains to be seen how to rigorously dene such clusters and what the mechanism is for these clusters to inuence transport properties. In contrast, another type of antibody (MAbG) was shown to exhibit elevated viscosity 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

with addition of more salt to solution, 13 indicating that antibody self-association could have a variety of mechanisms, likely hydrophobic interactions in the case of MAbG. Neutron scattering experiments have indicated that the long-range electrostatic repulsions were screened and it was proposed that the antibodies clustered hierarchically through the exposure of localized sites with short-range attraction. 14 Although these fruitful insights into the physics of concentrated protein solutions represent progress, fundamental understanding and predictability are still lacking in this eld. Case-by-case investigations cannot handle the complexity and variety of possible protein self-association modes. Thus, models and computational tools are essential for a better understanding and prediction of bulk transport properties of concentrated protein solutions. Traditional MD simulation methodologies track the positions of all the atoms in molecules, and thus are useful when applied to a small number of protein molecules on short time scales. Past works have yielded important information on the detailed structure and dynamics of a single MAb molecule, 9,1517 as well as binding patterns between two MAbs to form a dimer. 10,13 However, the computational power required for an all-atom MD simulation of larger systems and longer time scales needed to sample the collective behavior and bulk properties of protein solutions is still unavailable. Simulations consisting of hundreds or thousands of molecules are necessary to capture the characteristics of clusters with comparable sizes in concentrated solutions, so a certain level of simplication is inevitable. Coarsegrained (CG) molecular dynamics simulations are widely used in the eld of biophysics. 1821 The complexity of macromolecules can be eectively reduced by grouping atoms into a single quasi-particle. The level of resolution for coarse-graining needs to be carefully selected depending on the specic application. The model can still be quite detailed, with several beads for every single residue, as applied to protein-protein binding, 22 aggregate formation, 23 and biomembrane 24 simulations. Alternatively, every bead can represent one or more amino acids, allowing more molecules to be simulated in the system given limited computing time and capacity. In this approach, the protein under investigation rstly requires a thorough 4

ACS Paragon Plus Environment

Page 4 of 41

Page 5 of 41 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

structural 25 or dynamic 26 analysis in order to identify the rigid domains of the molecule. It is essential that the CG site mapping reproduce either critical structures or dynamical modes within the molecule. 19 Then, the interactions between the CG sites can be identied and parameterized. 27

MD Brownian Yukawa vdW HI

CG Figure 1: Comparison between molecular dynamics simulation with explicit and coarsegrained simulation with implicit solvent. The interactions between protein and solvent molecules have to be included in form of coarse-grained forces, such as the Brownian stochastic force and hydrodynamic forces. Screening eect due to the solvent also changes the strength of electrostatic and van der Waals interactions. As shown in gure 1, an important reduction in the complexity level when transitioning from all-atom MD to CG simulation is application of an implicit instead of explicit solvent. 20 When applying an implicit solvent model, special considerations have to be made for the form of interactions between CG particles. Among those is a proper account for the solution thermostat in the form of the Brownian stochastic force acting on the coarse-grained elements and the corresponding hydrodynamic force exerted on the same. In addition, the strength of electrostatic interactions and dispersion forces are weakened due to dielectric screening from the implicit solvent. These physical processes are not always considered, which limits accuracy and predictive capability of dynamic simulations. In the present work, we will emphasize the proper scaling of the interactions between CG elements, aiming at reproducing 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 41

experimental results for structure and dynamics in concentrated protein solutions through a rational framework for CG simulation of monoclonal antibodies and other macromolecular structures.

Theory and Methods Coarse-grained model

MAb 1

MAb 2

Figure 2: Coarse-grained 12-bead models of MAb1 and MAb2. 28,29 The molecular structures in the background are the crystal structure of human IgG 16 for comparison. Blue beads represent beads with positive charge and red ones are those with negative charge. The proprietary molecules MAb1 and MAb2 11,12,28 from Genentech are selected as the model monoclonal antibody molecules in this study. The Y-shaped antibody molecule has an anisotropic structure and possesses highly anisotropic interactions, caused by hydrophobic patchiness and a heterogeneous charge distribution. A previous 12-bead CG model 28,29 derived from all-atom molecular dynamics simulations is used in the present work to capture this anisotropy. As shown in gure 2, atoms are grouped into CG beads, which interact with each other by dierent types of interactions. Bond, angle, Urey-Bradley (UB), and dihedral interactions are considered. The potential forms of these interactions can be found in table 1. In order to obtain the spring constants and equilibrium values, published results on uctuations in all-atom MD simulations of a single protein were used. 28 The equilibrium bond length, req , angle, θeq , and dihedral, d, are determined by averaging the MD trajectory, and 6

ACS Paragon Plus Environment

Page 7 of 41 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 spring constants, k∗ , are evaluated through equipartition by computing the magnitude of dierent structural uctuations. 15,26 Table 1:

Potential forms of intramolecular interactions in the coarse-grained

model

Interaction Bond Angle Urey-Bradley (UB) Dihedral

Potential form Ebond = kbond (r − req )2 Eangle = kangle (θ − θeq )2 EU B = kU B (r − rU B )2 Edihedral = kdihedral [1 + cos(ϕ − d)]

In this CG model, charges on the CG domains are computed by summing up the partial charges of residues within each domain based on the CHARMM force eld. The values of charges on each CG bead were also determined in the prior work with this 12-bead CG model. 28 The colors of the beads in gure 2 represent the sign of these lumped charges. MAb1 has more heterogeneous charge distribution, indicating attraction between the molecules might be driven by dipolar interactions, while MAb2 has higher net charge and the surface charges are more uniformly distributed, indicating the molecule will aggregate less readily than MAb1.

Brownian Dynamics simulation with hydrodynamic interactions Treating globular proteins as colloidal particles has been proven to be an eective method for understanding their mesoscopic behavior. 27,30 In the present work, colloidal interactions are applied to each coarse-grained bead in the antibody. Brownian Dynamics (BD) simulation is used to integrate particle trajectories, for its eciency, simplicity, and ability to reproduce dynamical trends in aggregation and self-association. 31,32 For any CG particle, its trajectory satises the Langevin equation:

m

dU = FH + FB + FP , dt

7

ACS Paragon Plus Environment

(1)

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

where m is the mass of the particle, U is the translational velocity vector, and the force vectors F represent: (i) hydrodynamic force FH , caused by drag of the solvent on the particle; (ii) stochastic force FB , responsible for Brownian motion of the particles; and (iii) deterministic non-hydrodynamic force FP , dened as the negative gradient of interaction potential,

FP = −∇E , which includes eects from bonded interactions among CG elements of a single molecule, excluded volume, electrostatic attraction/repulsion and dispersion interactions among dierent molecules. For a CG bead in an antibody, the characteristic time scale of momentum relaxation  the inertial time scale  is τ I = m/ηs a, where a is the hydrodynamic radius of the bead (2 nm, listed in table 2) and ηs is the viscosity of the solvent. This time scale is approximately

1.2 × 10−11 s; however, the diusive time scale of the same bead is τ D = 6πηs a3 /kB T , where T is the temperature, is approximately 3.3 × 10−8 s at room temperature. Thus, dierent from all atom MD simulations, the inertial term in the equation of motion, m(dU/dt), can be neglected as the momentum relaxes much faster than all other dynamic processes. Hydrodynamic interactions are long-ranged and decay with the inverse of the distance between suspended objects. This reects the fundamental nature of the ow eld induced by motion of colloidal objects in a viscous uid. Present in any solution, hydrodynamic interactions are crucial to the dynamics of colloidal dispersions, and including them in coarse-grained simulations of colloidal processes has been shown to signicantly improve the predictability of kinetic phenomena such as aggregation and gelation. 3335 Enhanced collective motion of colloids due to hydrodynamic interactions changes the modes of relaxation in protein solutions as well, and simulations have shown their inclusion is necessary to describe the kinetics of protein association. 36 More importantly, hydrodynamic interactions are indispensable for quantitative evaluation of transport properties, 3739 such as sedimentation rate, diusivity, and viscosity, which are critical for the processing and administration of therapeutic proteins. Hydrodynamic interactions set the rate of diusive transport processes across large length scales in colloidal dispersions, and are the single most important source for viscous dissipation 8

ACS Paragon Plus Environment

Page 8 of 41

Page 9 of 41 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

in concentrated formulations. Their neglect in coarse-grained models can lead to incorrect predictions of transport properties. Application of these incorrect predictions in engineering cases will lead to failure in formulation processes, such as ultraltration/dialtration and ultracentrifugation. Thus, it is essential to include good models for the hydrodynamic interactions among suspended particles in CG models to obtain self-associated structures whose thermal relaxation processes and resistance to ow match well the physics in experiments. The hydrodynamic force FH in the over-damped regime is linear in the bead velocity:

FH = −R · U, where R is the hydrodynamic resistance tensor describing the hydrodynamic coupling between force and translation of all the particles. Neglecting inertia, the Langevin equation is:

0 = −R · U + FB + FP .

(2)

Inversion of resistance tensor results in the mobility tensor, M = R−1 , which can be used to convert the over-damped Langevin equation to a stochastic dierential equation for the particle positions x:

( ) dx = M · FB + FP . dt

(3)

Hydrodynamic and Brownian forces In the present work, the mobility tensor is approximated by the Rotne-Prager-Yamakawa (RPY) tensor MRP Y , which accounts for the far-eld translation-force coupling: 40

Y MRP αβ

   1  I  6πηs a   ( ) a2 2 1 = 1 + ∇ (I + ˆ rˆ r) 3 8πηs r     ] [    1 (1 − 9r )I + 3r ˆ r ˆ r 6πηs a 32a 32a

if α = β, if α ̸= β and r ≥ 2a,

(4)

if α ̸= β and r < 2a,

Y where MRP is the block at row α, column β of matrix M, and ˆ r and r are the unit vector αβ

and distance between the centers of particles α and β , respectively. According to the uctuation-dissipation theorem, the mean value and the autocorrelation 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 41

function of the Brownian stochastic force FB can be characterized by:

⟨FB ⟩ = 0, ⟨F (0)F (t)⟩ = 2kB T Rδ(t), B

(5)

B

where δ(t) is the Dirac delta function. Therefore, discretizing the force balance in time with the Euler-Maruyama integrator yields a discrete displacement for each of the CG elements: 32

x(t + ∆t) = x(t) + M · FP ∆t + kB T ∇ · M∆t + ∆xB ,

(6)

where x(t) is the position vector of the particles at time t. ∆xB is the stochastic displacement and satises ⟨∆xB ⟩ = 0 and ⟨∆xB ∆xB ⟩ = 2kB T M∆t. A high-performance algorithm aimed at Brownian Dynamics simulations, named the Positively Split Ewald (PSE) method, 41 was recently developed to vastly accelerate these calculations. The algorithm decomposes the hydrodynamic interactions into a local contribution and global contribution, which are evaluated independently using high performance algorithms. The utilized implementation of the PSE method was built as a plugin to HOOMDblue, 42,43 a molecular dynamics suite optimized for massively parallel processing on GPUs. The PSE method scales linearly in the number of suspended objects and to our knowledge is the fastest algorithm for coarse-grained Brownian Dynamics simulations with the RPY mobility. It is applied in the present work to simulate up to 4096 suspended antibodies.

Electrostatics The charge distribution on protein surfaces originates from the dierent dissociation tendency of individual residues. The electrostatic interactions between molecules can be calculated or measured through various methods. 44 For explicit solvent models, the charged surface of the protein can orient the solvent molecules to align with the electrostatic eld, generating an induced potential with opposite direction. Therefore, the interactions between two proteins 10

ACS Paragon Plus Environment

Page 11 of 41 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

are weakened due to the polarized solvent by a factor of the relative permittivity ϵr . For the implicit solvent model, this factor needs to be included when calculating electrostatic forces. Dissolved salt changes the strength of electrostatic interactions through a dierent mechanism called electrostatic screening, which weakens the electrostatic potential between CG beads. Ions in the solution tend to approach charged colloidal particles or protein surfaces with opposite charge. This will result in an accumulation of free charge surrounding the particle satisfying Boltzmann distribution:

nk = nkb exp(−ez k ψ/kB T ),

(7)

where nk and nkb are the number density of ion k locally and in the bulk. z k is the charge number, and ψ is electrostatic potential satisfying Poisson's equation ϵr ϵ0 ∇2 ψ = −ρ(f ) . ∑ k k (f ) Substituting the charge density gives the ion k ez n expression for the free charge ρ Poisson-Boltzmann equation:

ϵr ϵ0 ∇2 ψ = −e



z k nkb exp(−ez k ψ/kB T ).

(8)

ion k

If the rst order Taylor expansion is used for the exponential term to linearize the equation, the electrostatic potential ψ around a spherical particle with charge q can be solved as: 45

ψ=

q exp(κa) exp(−κr) , 4πϵr ϵ0 1 + κa r

(9)

where κ−1 is the Debye length and determined by ionic strength:

κ2 =

∑ ( )2 e2 nkb z k . ϵr ϵ0 kB T ion k

(10)

As for interactions between two charged particles with radii a1 and a2 , Bell, Levine, and McCartney 46 used linear superposition of each particle to approximate the overall potential,

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 41

and demonstrated that the electrostatic energy Eelec can be approximated as a form of the Yukawa potential:

( Eelec =

q1 q2 4πϵr ϵ0

)(

exp(κa1 ) 1 + κa1

)(

exp(κa2 ) 1 + κa2

)

exp(−κr) . r

(11)

This approximation gives less than 10% error as long as the distance r is longer than either radius, which suces for the purpose of our simulation. 47

Dispersion forces Dispersion forces are by their nature non-additive many-body interactions. 48 In addition, the expression and strength of dispersion interactions between coarse-grained particles are not the same as those between atoms (such as the Lennard-Jones potential). Instead, van der Waals interactions in the DLVO (Derjaguin-Landau-Verwey-Overbeek) framework, which is commonly used for colloidal dispersions, is a more appropriate force eld for CG simulation. 27 Hamaker 49 assumed additivity and integrated the interactions between each pair of molecules in two particles to approximate van der Waals interactions. The interaction EvdW is then:

EvdW

AH =− 2 π

∫ V1

∫ V2

dx1 dx2 , r6

(12)

where AH is Hamaker constant. For two spheres, the integral can be expressed as: 49

EvdW

[ ] 2a1 a2 2a1 a2 r2 − (a1 + a2 )2 AH =− + + ln 2 . 6 r2 − (a1 + a2 )2 r2 − (a1 − a2 )2 r − (a1 − a2 )2

(13)

When the two particles are of the same size a, it can be shown that this expression decays ( a )6 H asymptotically with − 16A in the long-range regime, which is the expected order for 9 r the dispersion interaction. AH can be directly related to the energy scale of intermolecular dispersion interactions. According to Lifshitz's continuum theory, 50,51 the Hamaker constant

12

ACS Paragon Plus Environment

Page 13 of 41 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

can be approximated as: 27,48,52,53

3 AH = kB T 4

(

ϵp − ϵs ϵp + ϵs

)2

[ ]2 ∞ ∑ 3 ϵp (iνn ) − ϵs (iνn ) + kB T , 2 ϵ (iν ) + ϵ (iν ) p n s n n=1,2,...

(14)

where ϵp/s and ϵp/s (iνn ) are static and frequency dependent permittivities of the protein or solvent. 52,54 For small protein molecules in water, the Hamaker constant is approximately

3kB T according to previous experimental measurements. 27,52 In the supplementary material, we present a sensitivity study exploring the eect of AH on structure. We nd that a smaller value of 0.4 kcal/mol gives the best agreement with experimental measurements of the structure factor. The determined value is smaller than previous experimental measurements because it measures the interaction strength between beads rather than whole protein molecules.

Results and Discussion Table 2: Parameters used in coarse-grained simulation

Parameter Hydrodynamic radius (a) Hard-sphere radius (aHS ) Relative permittivity (ϵr ) Hamaker constant (AH ) Temperature (T ) Time step (∆t) Simulation time (tS ) Number of antibodies (N ) Antibody concentration (C ) NaCl concentration (CNaCl )

Value 2 nm 1.5 nm 80 0.4 kcal/mol 300 K 2.17 ps 2.17 µs 512, 4096 10 - 200 mg/ml 15, 150 mM

Solutions of two therapeutic monoclonal antibodies MAb1 and MAb2 with various concentrations from 10 mg/ml to 200 mg/ml were simulated. Two representative NaCl concentrations of 15 mM and 150 mM were compared to illustrate the eects of a long or short Debye screening length (κ−1 ). Other simulation parameters are listed in table 2. It should 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

be noted that the hydrodynamic radius of each bead is larger than the hard-sphere radius, which is an artifact of coarse graining. The hydrodynamic radius a is determined to ensure the single molecule diusivity is recovered, while the hard-sphere radius aHS is determined by matching the experimentally measured structure factor. A description of the approach employed to determine the parameters a, aHS , and AH as well as a sensitivity analysis of the presented data to their values can be found in the supplementary material. The structure factor: S(q), short-time wave-vector dependent diusivity: DS (q), short-time self-diusivity:

DSS , and viscosity η are computed and compared with experimental results.

Structure Small-angle neutron scattering (SANS) is one of the experimental techniques applied to probe the solution structure and protein-protein interactions (PPI). 11,55,56 The structure factor S(q), which is proportional to the power spectrum of number density uctuations in the solution, can be easily obtained from SANS results which probes spatial correlations among atomic nuclei with large scattering cross sections. Low wave-vector, q , and high q regions of the structure factor represent long-range and short-range spatial correlations between suspended macro-molecules, respectively. The q → 0 limit of S(q) is proportional to the compressibility of the colloidal dispersion, and serves as an indicator of net attraction/repulsion. High values of S(0) indicate a more compressible microstructure. Readers with further interest in the interpretation of SANS results are referred to Ref. 11. Figure 3 depicts the calculated structure factor S(q) from CG simulation and corresponding experimental results from SANS. 11 Excellent agreement can be observed for all of the available experimental data, conrming the validity of the simulation method. At lower salt concentration (15 mM NaCl) where electrostatic interactions dominate, the dierent charge distributions between the two antibodies have great inuence on the solution structure. MAb1, which has a more heterogeneous charge distribution and lower net charge, has higher structure factor at low q . A distinctive increase of S(q) as q decreases for concen14

ACS Paragon Plus Environment

Page 14 of 41

Page 15 of 41

1.2 1.0

MAb1 150mM NaCl

MAb1 15mM NaCl

S(q)

0.8 0.6 Sim.

0.4

0.0 1.2 1.0

Expt.

10mg/ml 50mg/ml 100mg/ml 150mg/ml 200mg/ml

0.2

MAb2 150mM NaCl

MAb2 15mM NaCl

0.8

S(q)

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

0.6 0.4 0.2 0.0 1

1

qa

qa

Figure 3: Structure factor of MAb1 (top) and MAb2 (bottom) aqueous solutions under concentrations from 10 mg/ml to 200 mg/ml. Symbol color and shape represent simulations with various concentrations. Dierent screening eects from 15 mM NaCl and 150 mM NaCl are illustrated. Dashed lines are experimental values from SANS experiments. 11 Note: Complex buer composition was used in the experiments, so their ionic strength might not be accurately equivalent with 15 mM/150 mM NaCl.

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

tration of 50 mg/ml is observed in the simulation. This observation is reported to be due to the anisotropic electrostatic attraction between MAb1 molecules, and is not explained by isotropic liquid theories. 11 The replication of this structural feature indicates that the 12-bead model is able to capture structural features due to shape anisotropy and charge distribution. No upward trend of S(q) at low q can be found for the 10 mg/ml solution, potentially because the solution is too dilute for the antibodies to form clusters and display long-range correlations. It is also possible that some features of the structure might be revealed at lower q values, which the system size used in the simulations is unable to reveal. Examining the real space structure suggests this is not the case. However, future experimental investigations of such dilute antibody solutions would prove helpful to conrm our predictions and further understanding. In previous CG modeling results, 28,29 either dense clusters or strong network were observed, which would give rise to high peak at low q and could not replicate the experimental measurements from SANS. The formation of very strong bonds between molecules in previous works was due to neglecting the water permittivity of approximately 80 but using the vacuum value of 1 instead. This resulted in 80 times stronger electrostatic interactions. In the present work, correct values were used, thus no dense cluster or strong network are observed. Instead, as shown in gure 4, loosely connected clusters are prevalent in the MAb1 solution and the whole dispersion remains homogeneous. In contrast to MAb1, MAb2 carries higher net charge and forms fewer and smaller clusters of monomers at low salt concentration. The repulsion between MAb2 molecules decreases the value of the structure factor at lower q . An isotropic charged sphere model can predict this feature of the structure factor, 11 but also includes a strong nearest neighbor peak, which was not observed in SANS experiments or the simulations in the present work. This indicates that the far-eld structure is mainly controlled by the net charge while the nonspherical shape prevents the antibodies from packing compactly. Increasing the salt concentration enhances the decay of electrostatic interactions. The 16

ACS Paragon Plus Environment

Page 16 of 41

Page 17 of 41

MAb 1

MAb 2 (b)

(c)

(d)

100mg/ml

(a)

150mg/ml

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 4: Cluster identication of simulations at 100 mg/ml (a, b) and 150 mg/ml (c, d) under 15 mM NaCl condition. Each bead represents a CG particle. Monomers, oligomers with no more than 10 molecules, and larger clusters are colored white, blue, and red, respectively. Antibodies are marked connected if one of the beads of one antibody is closer than 1.75a with a bead of the other. The gures were generated in VMD. 57

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 5: Scatter plot showing the scaling relationship between radius of gyration Rg and size N of each identied cluster in MAb solutions with 15 mM NaCl. An N 1/2 scaling can be clearly observed for both MAb1 and MAb2 solutions. Debye length κ−1 is 2.5 nm with 15 mM NaCl, but decreases to 0.78 nm with 150 mM NaCl. As shown on the right in gure 3, neither the anisotropic attraction between MAb1 nor the repulsion between MAb2 is reected in S(q). The structures are nearly identical for the entire range of concentrations. With the electrostatic interactions screened, the eect of dispersion interactions is evident in the solutions structure. The S(q) at low q values is larger due to the attractive dispersion interactions, when compared with simulations accounting only for hard-sphere repulsions, shown in the supplementary material. The only available SANS data at 150 mg/ml MAb1 with 150 mM NaCl supports the simulation results. 12 In order to better understand the structure of simulated antibody solutions, cluster analysis is performed on systems with 15 mM NaCl. Two antibody molecules are identied as connected if the nearest distance between their beads is shorter than a cuto length of 1.75a. The cuto length is determined by computing the radial distribution function and locating the rst local minimum. A discussion and sensitivity analysis of this choice can be found in the supplementary material. We observe an abundance of large clusters that break and reform continually due to Brownian motion. As shown in gure 4 and 5, with 15 mM NaCl, many more clusters are identied in MAb1 solution than MAb2. This result is consistent 18

ACS Paragon Plus Environment

Page 18 of 41

Page 19 of 41

with the structure factor calculation. In gure 5, radius of gyration, Rg , of each antibody cluster is also computed and illustrated as a function of cluster size. An N 1/2 scaling is observed for all the cases, conrming the clusters are not compact but remain fractal and occupy more volume than a geometrically dense packing.

Dynamics and viscosity 150mM NaCl

15mM NaCl

6

DS(q) (Å2/ns)

5 4 3 MAb1

2 1

0 1.2

MAb2

10mg/ml 50mg/ml 100mg/ml 150mg/ml 200mg/ml

1.0 0.8

H(q)

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

0.6 0.4 0.2 0.0 1

1

qa

qa

Figure 6: Short-time wave-vector dependent diusivity DS (q) (top) and hydrodynamic function H(q) (bottom) of MAb1 (solid lines) and MAb2 (dashed lines) aqueous solutions under concentrations from 10 mg/ml to 200 mg/ml. Dierent screening eects from 15 mM NaCl (left) and 150 mM NaCl (right) are illustrated. Diusivity is a critical transport property of monoclonal antibodies inuencing both manufacture and drug delivery. It reects the dynamics and relaxation of a solution of Brownian 19

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 20 of 41

particles or macromolecules. Quantication of diusivity might sometime be ambiguous and misleading in literature because either collective diusivity or self-diusivity can be actually referred to with the blanket term diusivity. Collective diusivity, also called gradient diusivity or bulk diusivity, is dened as the ratio between diusive ux and concentration gradient in a bulk protein solution as used in Fick's law. This is a critical macroscopic transport property highly relevant to the performance of a ltration process. 58 Self-diusivity, or tracer diusivity, measures the diusion dynamics of a single macromolecule in motion among its neighbors, which is an important microscopic quantity inuencing the eciency of drug delivery. Collective diusivity and self-diusivity are the low q limit and high q limit of wave-vector dependent diusivity, D(q), respectively. In experiments, the wave-vector dependent diusivity is usually measured by scattering techniques, such as dynamic light scattering (DLS) and neutron spin echo (NSE). Intermediate scattering function (ISF) I(q, t), which describes the decay of number density uctutations in time, can be obtained from the autocorrelation of the scattering intensity in either of these experiments. 59 The decaying rate of the ISF at dierent q values characterizes relaxation processes across dierent length scales. In simulations, analogous calculations of the intermediate scattering function can be performed by directly analyzing particle trajectories. For protein centers of mass at time t denoted, yi (t), the intermediate scattering function can be written as:

1 I(q, t) = N



N ∑

⟩ −iq·(yj (t′ +t)−yk (t′ ))

e

,

(15)

j,k=1

where the angle brackets indicate an ensemble average over many congurations. In the short time limit, the intermediate scattering function I(q, t) decays exponentially and the time constant is determined by the wave-vector dependent diusivity DS (q):

I(q, t)/I(q) = e−q

2 D S (q)t

.

(16)

As shown in gure 6, short-time wave-vector dependent diusivity, DS (q), depends strongly 20

ACS Paragon Plus Environment

Page 21 of 41 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

on protein type and solution conditions. The hydrodynamic function H(q) = DS (q)S(q)/D0 is also obtained (in which D0 is the single antibody diusivity in free space). The low q limit of hydrodynamic function is proportional to the sedimentation velocity of protein in bulk solution. The simulated MAb1 and MAb2 have approximately equal DS (q) at higher

q values, regardless of salt concentration. This is because the RPY tensor gives similar individual dynamics in the short-time limit. However, dierent solution structures inuence collective dynamics signicantly in the lower q region. For example, the diverging DS (q) of MAb2 at low q and lower salt concentration occurs because of the strong electrostatic repulsion between the molecules. The MAb2 solution is highly incompressible. In contrast, the MAb1 solution is rather compressible and DS (q) reaches a constant value at low q . When the salt concentration is 150 mM, on the other hand, the two antibody solutions have similar diusivities, which only mildly increase in the low q region. The hydrodynamic function H(q) turns out to be similar for both antibody types and both salt concentrations, which means the variation in DS (q) is primarily due to dierences in the structure factor. The short-time self-diusivity, DSS , is the high q limit of short-time wave-vector dependent diusivity. In addition to utilizing intermediate scattering function (ISF), there are two other methods to calculate DSS . The rst method is by mean squared displacement (MSD):

⟨∆R2 ⟩ , ∆t→0 6∆t

DSS = lim

(17)

where ⟨∆R2 ⟩ is mean squared displacement on the protein center of mass within time interval

∆t. DSS can be easily obtained from molecular trajectories. The other method, which has been applied in the present work, is static in nature and relies on snapshots of the solution structure. In this approach, static structure of the antibody solution is sampled from dierent time steps of the simulation. During the calculation of every selected frame, each MAb molecule in the solution is considered rigid and the mobility correlation between translational velocity and force on these rigid entities is evaluated.

21

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 22 of 41

The RPY model for hydrodynamic interactions is augmented by the assumption of molecular rigidity which imposes additional constraints for the relationship between force and velocity in each macromolecule. The total force F and torque T on the whole antibody i is simply the summation of forces or moments exerted on each CG bead α in it:

Fi =



fα ,

α∈i

Ti =



(18)

Hi,α · fα ,

α∈i

where Hi,α is:

 Hi,α



0 −(zα − zi ) yα − yi      = z − z 0 −(x − x ) i α i ,  α   −(yα − yi ) xα − xi 0

(19)

and fα is force on bead α. The velocity u of each CG bead α satises:

uα = Ui + HTi,α · Ωi ,

(20)

where Ui and Ωi are the translational and angular velocity of the antibody i, respectively. Since the transformations in (18) and (20) are linear, we can easily dene a grand matrix

Σ that correlates force, torque, translational velocity, and angular velocity on each protein molecule to bead force and velocity:

  F  =Σ·f T

  U and u = ΣT ·   . Ω

(21)

The RPY model for bead-bead interactions is used: u = M · f , and we obtain a resistance relation for the total force/torque and translational/angular velocity of hydrodynamically

22

ACS Paragon Plus Environment

Page 23 of 41 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

interacting, rigid antibody molecules:

  F T −1  =Σ·M ·Σ T

  U ·  , Ω

(22)

with the matrix product Σ · M−1 · ΣT acting as the resistance tensor of the suspension of rigidied proteins. This product can be divided into blocks representing dierent pairs of coupling:

       F  RFU RFΩ  U  ·  .  = Ω T RTU RTΩ

(23)

The grand resistance tensor can be inverted to obtain the grand mobility tensor:

   −1 MUF MUT  RFU RFΩ   =  MΩF MΩT RTU RTΩ

(24)

and the short-time self-diusivity is the trace of mobility tensor MUF :

DSS =

kB T tr MUF , 3N

(25)

where N is the total number of antibodies. An algorithm has been developed to perform this calculation eciently and the full details can be found in Ref. 60. The dashed lines in gure 7 illustrate the computed short-time self-diusivity DSS using the methodology mentioned above and assuming each MAb molecule is rigid. A comparison with experimental values from NSE measurements is made as well. As expected, selfdiusivity from both experiments and simulations decreases with increasing concentration. The computed diusivity of MAb1 is slightly lower than that of MAb2 with 15 mM NaCl, which is consistent with NSE results. However, overall the simulated diusivity shares similar value regardless of antibody type and NaCl concentration. In previous SANS work, 11 equivalent volume fractions ϕ were obtained by tting the structure factor. The linear re23

ACS Paragon Plus Environment

The Journal of Physical Chemistry

15mM NaCl

1.0

150mM NaCl

0.8

DSS/D0

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 41

0.6 0.4 MAb1

MAb2

0.2 Expt. Single Cluster

0.0

0

50

100

150

0

50

C (mg/ml)

100

150

C (mg/ml)

Figure 7: Comparison of short-time self-diusivity DSS between simulation and experiment. D0 is the single antibody diusivity in free space. Open symbols represent experimental values. Dashed and solid lines represent the computed diusivity assuming each antibody or each identied cluster to be rigid, respectively. The blue dash-dotted lines represent the theoretical value of random hard-sphere suspension diusivity. 61 lation for random hard spheres 61 DSS /D0 = 1 − 1.83ϕ is recovered if using that equivalent volume fraction, shown by the blue dash-dotted lines. The recovery of correct equivalent volume fraction demonstrates our model can successfully describe the dispersion microstructure and individual dynamics of each MAb molecule. However, the lower experimental diusivity indicates there are missing constraints between interacting molecules not considered in the computation, which will be discussed in detail in the next section. The viscosity of therapeutic monoclonal antibodies is a key parameter for processing and administration. However, prediction via theory or computational tools has proven challenging. The viscosity in a dispersion of particles consists of contributions from distinct sources: the solvent, hydrodynamic interactions, inter-particle interactions between the coarse-grained beads, and Brownian motion of the molecules. The bulk stress, σ in a dispersion deformed at a xed rate of strain, e, is given by: 62

σ = −pf I + 2ηs e + σ H + σ P + σ B ,

24

ACS Paragon Plus Environment

(26)

Page 25 of 41 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

where pf is the pressure of solvent, and e is the rate-of-strain tensor. The term 2ηs e is the contribution to the stress from the solvent. The last three terms in (26) are the particle phase contributions to the stress: σ H is the hydrodynamic contribution from drag of the freely suspended bodies in shear ow; σ P is the contribution from deterministic inter-particle interactions including bonded interactions, excluded volume, electrostatic interactions and dispersion interactions; σ B is the contribution arising from Brownian relaxation of the dispersion. With the RPY approximation, direct coupling of hydrodynamic stresses to particle motion is neglected. The hydrodynamic contribution can be simplied to the Einstein's expression σ H = 5ηs ϕe, where ϕ is the volume fraction of the protein molecules in solution. Similarly, the RPY approximation necessitates that the Brownian contribution can be simplied to the isotropic pressure contributed by the thermal energy: σ B = −nkB T I, where

n is the number density of protein molecules. With this same model, the contribution from deterministic interactions is then the only source of dierences between MAb1 and MAb2. This contribution can be expressed as:

σ P = −n⟨xFP ⟩.

(27)

According to linear response theory, the zero-shear viscosity of a dispersion can be calculated by performing equilibrium simulations and integrating the time-dependent stress autocorrelation function: 63

η=

V kB T





⟨σxy (0)σxy (t)⟩dt,

(28)

0

where V is the volume of the simulation box. In gure 8, the xFP contribution to zero-shear viscosity is plotted against the solution concentration. The viscosity does not signicantly depend on either antibody type or salt concentration. We nd that the xFP contribution to the viscosity of MAb2 is slightly higher than to MAb1 at 100 mg/ml and 150 mg/ml, but lower at 200 mg/ml. In experimental studies, however, the viscosity of both MAb solutions 25

ACS Paragon Plus Environment

The Journal of Physical Chemistry

grow far more rapidly with concentration than in the simulations, 6,9,12,13 as shown by the dashed gray and light red lines in gure 10. At 150 mg/ml, MAb1 should have a viscosity more than 100 times that of water and 10 times large than the viscosity of MAb2. NaCl 15mM 150mM MAb1 MAb2

2.0

/

S

1.5 xF 0

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 41

1.0

0.5

0.0 0

50

100 150 C (mg/ml)

200

Figure 8: Zero-shear viscosity η0xF contributed by non-hydrodynamic interactions of MAb1 and MAb2 solutions as a function of antibody concentration. The viscosity is computed through Green-Kubo relation. Viscosity measures the energy dissipation at a certain deformation rate. The underestimation of solution viscosity in simulations suggests that the actual relaxation occurs more slowly than predicted by simulations based on the RPY model. Much like the short-time diusivity, the rates of energy dissipation due to relative motion within the dispersion are underestimated by our approach in spite of an accurately resolved microstructure on the length scale close to the molecular size. Resolution of this discrepancy requires accounting for the constraints and resulting missing degrees of freedom in the coarse-grained model.

Eects from intermolecular constraints We hypothesize that the discrepancy in self-diusivity and viscosity between experiments and our simulations applying the RPY model originates from the lack of strong and slowly relaxing coupling between nearly touching proteins. This coupling could have many sources that are neglected in the current course-grained model. We discuss two below and propose 26

ACS Paragon Plus Environment

Page 27 of 41

a method of approximation to account for these eects without increasing complexity.

15

r

10

R21 /6 !sa FU

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

RPY

5

Refined

xx yy zz

0 2

4

6

r/a Figure 9: Force-translation coupling R21 FU between nearly touching antibodies. Solid and dashed lines represent the 12-bead RPY model and a rened composite-bead model, respectively. Dierent colors stand for the three diagonal elements of the resistance matrix. The conguration of the two antibodies is shown in the image as well. The denition of the axes based on the left antibody is illustrated. The blue and red beads (positive and negative charges) represent the beads in the 12-bead model, and the smaller white transparent ones are the particles of the rened model. One potential source is the hydrodynamic coupling between protein molecules. Figure 9 depicts the hydrodynamic force resisting relative motion of two MAb molecules as a function of the distance between the Fab domains. Here the molecules are held in a particular orientation and one is made to translate in three orthogonal directions to the Fab domain on the other. The molecular shapes are frozen while dierent coarse-grained models for resolving their hydrodynamic interactions are compared. The resistance tensor denoted R21 FU represents the ratio of the measured force to the applied velocity. For rigid, impermeable bodies with nite curvature, this quantity should diverge near contact due to lubrication ows. These strong forces slow the relative motion between nearly touching bodies. The eect of strong hydrodynamic coupling is to freeze out relative motion but in favor of collective motion 27

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

of the bodies. Although lubrication in its continuum scale form is not likely to be active among monoclonal antibodies, we show that decomposing the 12-bead model into a higher order composite-bead model leads to a signicant increase in the predicted hydrodynamic resistance over the more naive RPY hydrodynamic model. Another source of dynamic constraint between nearly touching proteins is the heterogeneous and deformable surfaces of the macromolecules. These are decorated with locally patchy electrostatic and hydrophobic moieties that could play an important role in sustaining the physical bonds between molecules. This comes in two modes neglected in the 12-bead coarse-grained model: additional friction for relative motion between contacting proteins, and the ability of physical bonds to sustain bending moments. Recent work 64 has shown that the 12-bead model provides a good estimate of the second virial coecient, so these localized interactions have no role in the global structure of the dispersion. Their only role rheologically may be to distribute stress across self-associated clusters more eciently. To model generally the inuence of these local constraints on dynamics and rheology, each MAb molecule within a self-associated cluster as shown in gure 4 is forced to satisfy an additional rigidity constraint accounting for these missing degrees of freedom in the coarse-graining process. The hypothesis is that the short-time relaxation of MAb clusters is dictated by collective motion of the cluster and not the motion of individual MAbs within it. By rigidifying the clusters in simulations, new calculations of the self-diusivity and the hydrodynamic contribution to viscosity (which is also called high-frequency viscosity) can be made and compared with calculations with only individual antibodies rigidied. Thus, the impact of cluster level collective relaxation modes and frozen local particle motions can be studied. Since the static structure factor in the simulations agrees well with the experimentally measured microstructure, we expect short-time self-diusivities and viscosities coming from simulations with and without this rigidity constraint and their comparison to experimental results to be a direct test of the hypothesis. Indeed, as the solid lines in gure 7 show, insisting on rigid, collective motion of self28

ACS Paragon Plus Environment

Page 28 of 41

Page 29 of 41 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

associated clusters in the simulations successfully distinguishes self-diusivities of MAb1 and MAb2 solutions with 15 mM NaCl. This distinction is not captured from DS (q) or

DSS measurements without the additional rigidity constraint. This indicates that a bias to collective cluster motion is a probable mechanism yielding dierent self-diusivities between MAb1 and MAb2. The rigid cluster constraint also reduces the self-diusivity to levels lower than experimental observations across the whole range of protein concentrations and salt concentrations. The absolutely rigid motion of clusters is likely too strict a constraint and leads too much viscous dissipation on short time scales. Consequently, there is an underprediction of the diusivity. However, the experimental values for both proteins are cleanly bounded by the two proposed approximations. Einstein's contribution to the stress resulting from dilute suspended solids: σ H = 5ηs ϕe, provides a poor estimate of the hydrodynamic contribution to viscosity among the constrained rigid clusters, as the solution is not dilute. Instead, a similar method to the self-diusivity calculation is applied to compute σ H directly. Multiple snapshots of the simulation are selected as samples of representative antibody congurations, and either each MAb molecule or each self-associated cluster can be assumed to move rigidly. The stress resulting from such motion in an imposed shear ow is used to compute the viscosity and to illustrate the inuence of rigidity constraints among MAbs on σ H . During the stress calculation, a ctitious strain rate e is imposed on the static structure of antibody solution. The force F, torque T, and stresslet S response on each rigid objects is linear in this strain rate with the relation given by the corresponding resistance tensors:

F = RFE · e, T = RTE · e,

(29)

S = RSE · e. The force and torque on the rigid bodies give rise to translational and rotational motion, as shown in (23). This motion also contributes to the hydrodynamic stress: σ H linearly 29

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 30 of 41

through the stresslet coupling resistance tensors RSU and RSΩ , so the total stress contributed hydrodynamically is: 65

1 ∑ (RSU · U + RSΩ · Ω − RSE · e) V N   −1    [ ] 1 ∑  RFE  RFU RFΩ  =−  − RSE  · e.  ·  RSU RSΩ ·  V N RTE RTU RTΩ

σH = −

(30)

In (30), 5 resistance tensors RFE , RTE , RSE , RSU , and RSΩ are still unknown, but can be easily evaluated given the rigid body congurations using the same approach applied in deriving (22). For a rigid body i consisting of beads α, its stresslet S is:

] 1 ∑[ T T (xα − xi ) fα + fα (xα − xi ) . Si = 2 α∈i

(31)

Similarly, an extra term for ane motion should be added to the velocity of each bead if the body is immersed in an imposed linear ow with strain rate tensor e:

uα = Ui + HTi,α · Ωi + e · (xα − xi ) .

(32)

Both of these relations are linear, so we can use tensors K and K′ to represent compactly these relations for all the beads:

S=K·f

  U and u = ΣT ·   + K′ · e, Ω

30

ACS Paragon Plus Environment

(33)

Page 31 of 41 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

and obtain the resistance relations:





RFE  −1 ′  =Σ·M ·K, RTE [ RSU

RSE = K · M−1 · K′ , ] T −1 RSΩ = K · M · Σ .

(34)

The total viscosity is evaluated by adding the hydrodynamic contribution and the interparticle interaction contribution. Figure 10 shows the comparison between viscosities computed with and without the rigid cluster assumption. The dashed lines in gure 10 show the total viscosity predicted when assuming each antibody is rigid with no rigid constraint among clusters. Similar to the self-diusivity without cluster rigidity, the predicted viscosity underestimates the experimental viscous dissipation and MAb1 appears only slightly more viscous than MAb2. The solid lines in the same gure show viscosity predictions with the rigidity constraint imposed on all self-associated clusters. A dramatic increase in viscosity can be observed for both MAb1 and MAb2. The predicted MAb2 viscosity exhibits quantitative agreement with the experimental measurements. The viscosity of MAb1 is much higher than MAb2, since the MAb1 molecules form larger self-associated structures in concentrated solution, as shown in gure 4. Despite the improvement of our prediction trends, however, the MAb1 viscosity computed still under-predicts the experimental value, indicating there may be other factors critical for determining protein solution viscosity. In summary, transport properties can be estimated by coarse-grained simulations of the equilibrium microstructure coupled to more sophisticated hydrodynamic calculations for transport properties that account in an approximate way for the missing constraints on localized motion. Without these constraints, transport properties predicted are not able to recover the dierences between two antibody solutions with high sequence similarity but distinct solution microstructure. The methodology applied herein may prove useful for both

31

ACS Paragon Plus Environment

The Journal of Physical Chemistry

MAb1

100

MAb2

Expt. Single

S

Cluster

/

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 32 of 41

10

1 0

20

40

60

80

100

120

140

160

C (mg/ml)

Figure 10: Viscosity of MAb1 and MAb2 solutions computed with rigidity constraint assumption compared with experiments. 12 Dashed and solid lines represent the computed viscosity assuming each antibody or each identied cluster to be rigid, respectively. The total viscosity is evaluated by adding the hydrodynamic contribution from static structure calculation and the inter-particle interaction contribution from Green-Kubo relation. rank ordering as well as quantitative prediction of protein solution transport properties.

Conclusions Structure, relaxation, and their roles in macroscopic transport properties of concentrated monoclonal antibody solutions were investigated through Brownian Dynamics simulations with hydrodynamic interactions. The structure factor, self-diusivity, and viscosity have been computed and agree with experimental measurements of the same. This is the rst coarse-grained model to reproduce the bulk transport properties of monoclonal antibody solutions using only microscopic parameters. The methodology adopted in the present work requires simulation of the equilibrium solution microstructure coupled to hydrodynamic transport property calculations invoking rigid constraints between bonded proteins. Such coarse-graining approaches may be extended to other protein and macromolecular systems as well. 32

ACS Paragon Plus Environment

Page 33 of 41 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 distinct anisotropic attraction of MAb1 molecules due to their charge distribution is successfully captured using a simple 12-bead protein model, conrming the important role of electrostatic interactions in antibody self-association. The anisotropic nature of the interaction makes it challenging to build theoretical models of the structure without specialized and potentially dubious parameter tting. Coarse-grained simulations seem a better mode of approximation. The hypothesis of rigid constraints within protein clusters is validated and found to be useful in predicting macroscopic behaviors of antibody solutions with low complexity models. The far-eld RPY tensor is sucient to predict the equilibrium self-associated structure of two MAbs, but over-predicts the self-diusivity and under-predicts the viscosity. When applying an additional constraint that all clustered molecules move rigidly, predicted results are improved signicantly with respect to experiments. This constraint has features of higher order hydrodynamic models meant to replicate missing degrees of freedom in the coarse-grained model. To improve the transport property prediction in computational models, future investigations should be focused on understanding the nature and mechanisms of constraints due to near eld protein-protein interactions and by improving computational frameworks for quantitative transport property prediction. For example, constructing a composite-bead model similar with that shown in gure 9 based on protein crystal structure could provide more detailed and rened description of hydrodynamics and localized interactions from heterogeneous protein surfaces. Near-eld hydrodynamic interactions can be modeled to desired precision with arbitrarily rened surface tessellation, 66 which will give appropriate resistance between the proteins instead of assuming overly rigid clusters. In addition, techniques to more accurately evaluate electrostatic interactions by solving Poisson-Boltzmann equation 67,68 can be incorporated in the framework to better describe near-eld electrostatic features and make it possible to quantify the inuence of such features on the dynamics of nearly touching macromolecules. 33

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 34 of 41

Supporting Information Sensitivity analysis of simulation parameters (hydrodynamic radius, hard-sphere radius and Hamaker constant). Sensitivity analysis of cuto length in cluster identication. Structure factor with only hard-sphere repulsion.

Acknowledgments This work was supported by Genentech Inc. and the MIT Portugal Program. The GPUs utilized in this study were in part donated by the NVIDIA Corporation. We gratefully thank P. Douglas Godfrin for useful discussions.

References (1) Keizer, R. J.; Huitema, A. D.; Schellens, J. H.; Beijnen, J. H. Clinical pharmacokinetics of therapeutic monoclonal antibodies. Clin. Pharmacokinet.

2010,

49, 493507.

(2) Shire, S. J.; Shahrokh, Z.; Liu, J. Challenges in the development of high protein concentration formulations. J. Pharm. Sci.

2004,

93, 13901402.

(3) Kanai, S.; Liu, J.; Patapo, T. W.; Shire, S. J. Reversible self-association of a concentrated monoclonal antibody solution mediated by Fab-Fab interaction that impacts solution viscosity. J. Pharm. Sci.

2008,

97, 42194227.

(4) Burckbuchler, V.; Mekhlou, G.; Giteau, A. P.; Grossiord, J. L.; Huille, S.; Agnely, F. Rheological and syringeability properties of highly concentrated human polyclonal immunoglobulin solutions. Eur. J. Pharm. Biopharm.

2010,

76, 351356.

(5) Goswami, S.; Wang, W.; Arakawa, T.; Ohtake, S. Developments and challenges for mAb-based therapeutics. Antibodies

2013,

2, 452500.

34

ACS Paragon Plus Environment

Page 35 of 41 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

(6) Liu, J.; Nguyen, M. D.H.; Andya, J. D.; Shire, S. J. Reversible self-association increases the viscosity of a concentrated monoclonal antibody in aqueous solution. J. Pharm. Sci. 2005,

94, 19281940.

(7) Hermeling, S.; Crommelin, D. J.; Schellekens, H.; Jiskoot, W. Structure-immunogenicity relationships of therapeutic proteins. Pharm. Res.

2004,

21, 897903.

(8) Sahin, E.; Grillo, A. O.; Perkins, M. D.; Roberts, C. J. Comparative eects of pH and ionic strength on protein-protein interactions, unfolding, and aggregation for IgG1 antibodies. J. Pharm. Sci.

2010,

99, 48304848.

(9) Yadav, S.; Laue, T. M.; Kalonia, D. S.; Singh, S. N.; Shire, S. J. The inuence of charge distribution on self-association and viscosity behavior of monoclonal antibody solutions. Mol. Pharmaceutics 2012,

9, 791802.

(10) Lapelosa, M.; Patapo, T. W.; Zarraga, I. E. Molecular simulations of the pairwise interaction of monoclonal antibodies. J. Phys. Chem. B

2014,

118, 1313213141.

(11) Yearley, E. J.; Zarraga, I. E.; Shire, S. J.; Scherer, T. M.; Gokarn, Y.; Wagner, N. J.; Liu, Y. Small-angle neutron scattering characterization of monoclonal antibody conformations and interactions at high concentrations. Biophys. J.

2013,

105, 720731.

(12) Yearley, E. J.; Godfrin, P. D.; Perevozchikova, T.; Zhang, H.; Falus, P.; Porcar, L.; Nagao, M.; Curtis, J. E.; Gawande, P.; Taing, R., et al. Observation of small cluster formation in concentrated monoclonal antibody solutions and its implications to solution viscosity. Biophys. J.

2014,

106, 17631770.

(13) Lilyestrom, W. G.; Yadav, S.; Shire, S. J.; Scherer, T. M. Monoclonal antibody selfassociation, cluster formation, and rheology at high concentrations. J. Phys. Chem. B 2013,

117, 63736384.

35

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 36 of 41

(14) Godfrin, P. D.; Zarraga, I. E.; Zarzar, J.; Porcar, L.; Falus, P.; Wagner, N. J.; Liu, Y. Eect of hierarchical cluster formation on the viscosity of concentrated monoclonal antibody formulations studied by neutron scattering. J. Phys. Chem. B

2016,

120, 278

291. (15) Brandt, J. P.; Patapo, T. W.; Aragon, S. R. Construction, MD simulation, and hydrodynamic validation of an all-atom model of a monoclonal IgG antibody. Biophys. J. 2010,

99, 905913.

(16) Saphire, E. O.; Parren, P. W.; Pantophlet, R.; Zwick, M. B.; Morris, G. M.; Rudd, P. M.; Dwek, R. A.; Staneld, R. L.; Burton, D. R.; Wilson, I. A. Crystal structure of a neutralizing human IgG against HIV-1: a template for vaccine design. Science 2001, 293, 11551159. (17) Lilyestrom, W. G.; Shire, S. J.; Scherer, T. M. Inuence of the cosolute environment on IgG solution structure analyzed by small-angle X-ray scattering. J. Phys. Chem. B 2012,

116, 96119618.

(18) Tozzini, V. Coarse-grained models for proteins. Curr. Opin. Struct. Biol.

2005,

15,

144150. (19) Saunders, M. G.; Voth, G. A. Coarse-graining methods for computational biology. Annu. Rev. Biophys. 2013,

42, 7393.

(20) Riniker, S.; Allison, J. R.; van Gunsteren, W. F. On developing coarse-grained models for biomolecular simulation: a review. Phys. Chem. Chem. Phys. 2012, 14, 1242312430. (21) Corbett, D.; Hebditch, M.; Keeling, R.; Ke, P.; Ekizoglou, S.; Sarangapani, P.; Pathak, J.; Van Der Walle, C. F.; Uddin, S.; Baldock, C., et al. Coarse-grained modeling of antibodies from small-angle scattering proles. J. Phys. Chem. B

36

ACS Paragon Plus Environment

2017,

121, 82768290.

Page 37 of 41 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

(22) Zacharias, M. Protein-protein docking with a reduced protein model accounting for side-chain exibility. Protein Sci.

2003,

12, 12711282.

(23) Bereau, T.; Deserno, M. Generic coarse-grained model for protein folding and aggregation. J. Chem. Phys.

2009,

130, 235106.

(24) Monticelli, L.; Kandasamy, S. K.; Periole, X.; Larson, R. G.; Tieleman, D. P.; Marrink, S. J. The MARTINI coarse-grained force eld: extension to proteins. J. Chem. Theory and Comput. 2008,

4, 819834.

(25) Arkhipov, A.; Freddolino, P. L.; Schulten, K. Stability and dynamics of virus capsids described by coarse-grained modeling. Structure

2006,

14, 17671777.

(26) Zhang, Z.; Pfaendtner, J.; Grafmüller, A.; Voth, G. A. Dening coarse-grained representations of large biomolecules and biomolecular complexes from elastic network models. Biophys. J. 2009,

97, 23272337.

(27) Arzensek, D.; Kuzman, D.; Podgornik, R. Colloidal interactions between monoclonal antibodies in aqueous solutions. J. Colloid Interface Sci.

2012,

384, 207216.

(28) Chaudhri, A.; Zarraga, I. E.; Kamerzell, T. J.; Brandt, J. P.; Patapo, T. W.; Shire, S. J.; Voth, G. A. Coarse-grained modeling of the self-association of therapeutic monoclonal antibodies. J. Phys. Chem. B

2012,

116, 80458057.

(29) Chaudhri, A.; Zarraga, I. E.; Yadav, S.; Patapo, T. W.; Shire, S. J.; Voth, G. A. The role of amino acid sequence in the self-association of therapeutic monoclonal antibodies: insights from coarse-grained modeling. J. Phys. Chem. B

2013,

117, 12691279.

(30) Leckband, D.; Sivasankar, S. Forces controlling protein interactions: theory and experiment. Colloids Surf. B

1999,

14, 8397.

(31) Heyes, D. M.; Melrose, J. R. Brownian dynamics simulations of model hard-sphere suspensions. J. Non-Newtonian Fluid Mech. 37

1993,

46, 128.

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 38 of 41

(32) Ermak, D. L.; McCammon, J. A. Brownian dynamics with hydrodynamic interactions. J. Chem. Phys. 1978,

69, 13521360.

(33) Cao, X. J.; Cummins, H. Z.; Morris, J. F. Hydrodynamic and interparticle potential eects on aggregation of colloidal particles. J. Colloid Interface Sci.

2012,

368, 8696.

(34) Varga, Z.; Wang, G.; Swan, J. W. The hydrodynamics of colloidal gelation. Soft Matter 2015,

11, 90099019.

(35) Varga, Z.; Swan, J. W. Hydrodynamic interactions enhance gelation in dispersions of colloids with short-ranged attraction and long-ranged repulsion. Soft Matter

2016,

12,

76707681. (36) Frembgen-Kesner, T.; Elcock, A. H. Striking eects of hydrodynamic interactions on the simulated diusion and folding of proteins. J. Chem. Theory and Comput.

2009,

5,

242256. (37) Brady, J. F.; Durlofsky, L. J. The sedimentation rate of disordered suspensions. Phys. Fluids 1988,

31, 717727.

(38) Ladd, A. J. C. Hydrodynamic transport coecients of random dispersions of hard spheres. J. Chem. Phys.

1990,

93, 34843494.

(39) Phillips, R. J.; Brady, J. F.; Bossis, G. Hydrodynamic transport properties of hardsphere dispersions. I. Suspensions of freely mobile particles. Phys. Fluids 1988, 31, 3462 3472. (40) Rotne, J.; Prager, S. Variational treatment of hydrodynamic interaction in polymers. J. Chem. Phys. 1969,

50, 48314837.

(41) Fiore, A. M.; Balboa Usabiaga, F.; Donev, A.; Swan, J. W. Rapid sampling of stochastic displacements in Brownian dynamics simulations. J. Chem. Phys.

38

ACS Paragon Plus Environment

2017,

146, 124116.

Page 39 of 41 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

(42) Anderson, J. A.; Lorenz, C. D.; Travesset, A. General purpose molecular dynamics simulations fully implemented on graphics processing units. J. Comput. Phys. 2008, 227, 53425359. (43) HOOMD-blue. https://codeblue.umich.edu/hoomd-blue.

2018.

(44) Gitlin, I.; Carbeck, J. D.; Whitesides, G. M. Why are proteins charged? Networks of charge-charge interactions in proteins measured by charge ladders and capillary electrophoresis. Angew. Chem. Int. Ed.

2006,

45, 30223060.

(45) Debye, P.; Hückel, E. On the theory of electrolytes. I. Freezing point depression and related phenomena. Physikalische Zeitschrift

1923,

24, 185206.

(46) Bell, G. M.; Levine, S.; McCartney, L. N. Approximate methods of determining the double-layer free energy of interaction between two charged colloidal spheres. J. Colloid Interface Sci. 1970,

33, 335359.

(47) Russel, W. B.; Saville, D. A.; Schowalter, W. R. Colloidal dispersions; Cambridge University Press: Cambridge, U.K.,

1989

(48) Israelachvili, J. N. Intermolecular and surface forces (Third edition); Academic Press: Cambridge, MA, U.S.A.,

2011

(49) Hamaker, H. C. The London-van der Waals attraction between spherical particles. Physica 1937,

4, 10581072.

(50) Lifshitz, E. M. The theory of molecular attractive forces between solids. Soviet Physics 1956,

2, 7383.

(51) Dzyaloshinskii, I. E.; Lifshitz, E. M.; Pitaevskii, L. P. General theory of van der Waals' forces. Soviet Physics Uspekhi

1961,

73, 153176.

(52) Roth, C. M.; Neal, B. L.; Lenho, A. M. Van der Waals interactions involving proteins. Biophys. J. 1996,

70, 977987. 39

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 40 of 41

(53) Bergström, L. Hamaker constants of inorganic materials. Adv. Colloid Interface Sci. 1997,

70, 125169.

(54) Hough, D. B.; White, L. R. The calculation of Hamaker constants from Liftshitz theory with applications to wetting phenomena. Adv. Colloid Interface Sci.

1980,

14, 341.

(55) Clark, N. J.; Zhang, H.; Krueger, S.; Lee, H. J.; Ketchem, R. R.; Kerwin, B.; Kanapuram, S. R.; Treuheit, M. J.; McAuley, A.; Curtis, J. E. Small-angle neutron scattering study of a monoclonal antibody using free-energy constraints. J. Phys. Chem. B

2013,

117, 1402914038. (56) Kim, H. S.; Martel, A.; Girard, E.; Moulin, M.; Härtlein, M.; Madern, D.; Blackledge, M.; Franzetti, B.; Gabel, F. SAXS/SANS on supercharged proteins reveals residue-specic modications of the hydration shell. Biophys. J.

2016,

110, 21852194.

(57) Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual Molecular Dynamics. J. Molec. Graphics 1996,

14, 3338.

(58) Belfort, G.; Davis, R. H.; Zydney, A. L. The behavior of suspensions and macromolecular solutions in crossow microltration. J. Membr. Sci.

1994,

96, 158.

(59) Berne, B. J.; Pecora, R. Dynamic light scattering: with applications to chemistry, biology, and physics;

Dover Publications: Mineola, NY, U.S.A.,

2000

(60) Swan, J. W.; Wang, G. Rapid calculation of hydrodynamic and transport properties in concentrated solutions of colloidal particles and macromolecules. Phys. Fluids

2016,

28,

011902. (61) Batchelor, G. K. Brownian diusion of particles with hydrodynamic interaction. J. Fluid Mech. 1976,

74, 129.

(62) Foss, D. R.; Brady, J. F. Structure, diusion and rheology of Brownian suspensions by Stokesian dynamics simulation. J. Fluid Mech. 40

2000,

407, 167200.

ACS Paragon Plus Environment

Page 41 of 41 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

(63) Frenkel, D.; Smit, B. Understanding molecular simulation: from algorithms to applications;

Academic Press: Cambridge, MA, U.S.A.,

2002

(64) Calero-Rubio, C.; Saluja, A.; Roberts, C. J. Coarse-grained antibody models for weak protein-protein interactions from low to high concentrations. J. Phys. Chem. B

2016,

120, 65926605. (65) Bossis, G.; Brady, J. F. The rheology of Brownian suspensions. J. Chem. Phys.

1989,

91, 18661874. (66) Delong, S.; Balboa Usabiaga, F.; Donev, A. Brownian dynamics of conned rigid bodies. J. Chem. Phys.

2015,

143, 144107.

(67) Lu, B.; Cheng, X.; Huang, J.; McCammon, J. A. An adaptive fast multipole boundary element method for Poisson-Boltzmann electrostatics. J. Chem. Theory Comput.

2009,

5, 16921699. (68) Yap, E. H.; Head-Gordon, T. New and ecient Poisson-Boltzmann solver for interaction of multiple proteins. J. Chem. Theory Comput.

2010,

6, 22142224.

MD Brownian Yukawa vdW HI

CG Figure 11: TOC graphic

41

ACS Paragon Plus Environment