Quantitative Prediction of Protein–Polyelectrolyte Binding Thermodynamics: Adsorption of Heparin-Analog Polysulfates to the SARS-CoV-2 Spike Protein RBDClick to copy article linkArticle link copied!
- Lenard NeanderLenard NeanderDepartment of Physics, Freie Universität Berlin, Arnimallee 14, Berlin 14195, GermanyInstitute of Chemistry and Biochemistry, Freie Universität Berlin, Takustraße 3, Berlin 14195, GermanyMore by Lenard Neander
- Cedric HannemannCedric HannemannDepartment of Physics, Freie Universität Berlin, Arnimallee 14, Berlin 14195, GermanyMore by Cedric Hannemann
- Roland R. Netz*Roland R. Netz*Email: [email protected]Department of Physics, Freie Universität Berlin, Arnimallee 14, Berlin 14195, GermanyMore by Roland R. Netz
- Anil Kumar Sahoo*Anil Kumar Sahoo*Email: [email protected]Department of Physics, Freie Universität Berlin, Arnimallee 14, Berlin 14195, GermanyMore by Anil Kumar Sahoo
Abstract
Interactions of polyelectrolytes (PEs) with proteins play a crucial role in numerous biological processes, such as the internalization of virus particles into host cells. Although docking, machine learning methods, and molecular dynamics (MD) simulations are utilized to estimate binding poses and binding free energies of small-molecule drugs to proteins, quantitative prediction of the binding thermodynamics of PE-based drugs presents a significant obstacle in computer-aided drug design. This is due to the sluggish dynamics of PEs caused by their size and strong charge–charge correlations. In this paper, we introduce advanced sampling methods based on a force-spectroscopy setup and theoretical modeling to overcome this barrier. We exemplify our method with explicit solvent all-atom MD simulations of the interactions between anionic PEs that show antiviral properties, namely heparin and linear polyglycerol sulfate (LPGS), and the SARS-CoV-2 spike protein receptor binding domain (RBD). Our prediction for the binding free-energy of LPGS to the wild-type RBD matches experimentally measured dissociation constants within thermal energy, kBT, and correctly reproduces the experimental PE-length dependence. We find that LPGS binds to the Delta-variant RBD with an additional free-energy gain of 2.4 kBT, compared to the wild-type RBD, due to the additional presence of two mutated cationic residues contributing to the electrostatic energy gain. We show that the LPGS–RBD binding is solvent dominated and enthalpy driven, though with a large entropy–enthalpy compensation. Our method is applicable to general polymer adsorption phenomena and predicts precise binding free energies and reconfigurational friction as needed for drug and drug-delivery design.
This publication is licensed under
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
Introduction
Results and Discussion
Polymer Desorption Free Energy
Dissociation Constant and Standard Binding Free Energy
contributions | WT | Delta |
---|---|---|
ΔF | 5.59 ± 0.63 kBT | 7.46 ± 1.17 kBT |
ΔFstretch | 2.06 ± 0.29 kBT | 2.51 ± 0.41 kBT |
Vb | 11.32 ± 2.02 nm3 | 11.89 ± 4.91 nm3 |
kBT ln (Vb/ V0) | 1.92 ± 0.11 kBT | 1.97 ± 0.25 kBT |
ΔFb0 | –9.57 ± 0.72 kBT | –11.94 ± 1.30 kBT |
Nexp | ΔFb0 (exp.)b | ΔFb0 (theory)b | KD (exp.)c | KD (theory)c |
---|---|---|---|---|
110 (Page et al.) | –11.11 ± 0.37 | –11.87 ± 0.72 | 15.0 ± 5.5 | 7.0 ± 5.0 |
274 (Nie et al.) | –12.17 ± 0.69 | –12.78 ± 0.72 | 5.2 ± 3.6 | 2.8 ± 2.0 |
Experimental values are reproduced from publications by Nie et al., ref. (15) (Available under a CC BY-NC 4.0 license. Copyright 2021 The Authors.) and Page et al., ref. (35) (Available under a CC BY-NC 4.0 license. Copyright 2023 The Authors).
Standard binding free-energy, ΔFb0 in kBT.
Dissociation constant, KD in μM.
Enthalpy–Entropy Decomposition
Relaxation Time for Binding of Charged Groups
Protein Conformational Transitions
Conclusions
Methods
MD Simulations
Models, Parameters, and Simulation Set-Up
Glycan-Conjugated RBD
Dynamic and Static Pulling Simulations
Umbrella Sampling Simulations
Constant-Force Stretching of LPGS
Simulation Data Analysis
iPFRC Model and Stretching Free Energy of LPGS
parameter | value |
---|---|
a0 | 366.82 ± 0.67 pm |
aKuhn | 0.873 ± 0.033 nm |
c | 0.793 ± 0.018 |
γ1 | 96 ± 36 nN |
γ2 | 500 ± 2000 nN |
Internal Energy Decomposition
Distance Criteria for Close Contacts and Bound Water and Ions
Root-Mean-Square Deviation
Root-Mean-Square Fluctuation of the RBD Structure
Relaxation Time for the RBD’s Loop Region Movement
Error Estimations
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacsau.4c00886.
Derivation of the standard free-energy of binding from a polymer desorption free-energy profile; additional figures (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgments
We acknowledge support provided by Deutsche Forschungsgemeinschaft Grant No. IRTG-2662 Project No. 434130070 “Charging into the future,” and by the European Research Council under the European Union’s Horizon 2020 research and innovation program Grant Agreement No. 835117. We gratefully acknowledge computing time on the HPC clusters at the Physics department, Freie Universität Berlin.
References
This article references 84 other publications.
- 1Kornberg, R. D.; Lorch, Y. Twenty-five years of the nucleosome, fundamental particle of the eukaryote chromosome. Cell 1999, 98, 285– 294, DOI: 10.1016/S0092-8674(00)81958-3Google Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXlt1Cgt7o%253D&md5=7996cade955faa759d9bca55759eb513Twenty-five years of the nucleosome, fundamental particle of the eukaryote chromosomeKornberg, Roger D.; Lorch, YahliCell (Cambridge, Massachusetts) (1999), 98 (3), 285-294CODEN: CELLB5; ISSN:0092-8674. (Cell Press)A review with ∼100 refs., including sections entitled the histones, the nucleosome, structure of the nucleosome, structure of chromatin fibers, nucleosomes repress transcription, histone acetylation, histone deacetylation, chromatin remodeling silencing, nucleosomes and transcription elongation, multiple mechanisms of activation and repression, viral infection and cancer, general principles, and perspectives.
- 2Kunze, K.-K.; Netz, R. Complexes of semiflexible polyelectrolytes and charged spheres as models for salt-modulated nucleosomal structures. Phys. Rev. E 2002, 66, 011918 DOI: 10.1103/PhysRevE.66.011918Google ScholarThere is no corresponding record for this reference.
- 3Parsaeian, A.; De La Cruz, M. O.; Marko, J. F. Binding-rebinding dynamics of proteins interacting nonspecifically with a long DNA molecule. Phys. Rev. E 2013, 88, 040703 DOI: 10.1103/PhysRevE.88.040703Google ScholarThere is no corresponding record for this reference.
- 4Shin, Y.; Brangwynne, C. P. Liquid phase condensation in cell physiology and disease. Science 2017, 357, eaaf4382 DOI: 10.1126/science.aaf4382Google ScholarThere is no corresponding record for this reference.
- 5Goedert, M.; Jakes, R.; Spillantini, M.; Hasegawa, M.; Smith, M.; Crowther, R. Assembly of microtubule-associated protein tau into Alzheimer-like filaments induced by sulphated glycosaminoglycans. Nature 1996, 383, 550– 553, DOI: 10.1038/383550a0Google Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28Xmt1GqsL8%253D&md5=8c572e7db12bc47fa3fa8be625421fa8Assembly of microtubule-associated protein tau into Alzheimer-like filaments induced by sulfated glycosaminoglycansGoedert, M.; Jakes, R.; Spillantini, M. G.; Hasegawa, M.; Smith, M. J.; Crowther, R. A.Nature (London) (1996), 383 (6600), 550-553CODEN: NATUAS; ISSN:0028-0836. (Macmillan Magazines)The paired helical filament (PHF) is the major component of the neurofibrillary deposits that form a defining neuropathol. characteristic of Alzheimer's disease. PHFs are composed of microtubule-assocd. protein tau, in a hyperphosphorylated state. Hyperphosphorylation of tau results in its inability to bind to microtubules and is believed to precede PHF assembly. However, it is unclear whether hyperphosphorylation of tau is either necessary or sufficient for PHF formation. Here we show that non-phosphorylated recombinant tau isoforms with three microtubule-binding repeats form paired helical-like filaments under physiol. conditions in vitro, when incubated with sulfated glycosaminoglycans such as heparin or heparan sulfate. Furthermore, heparin prevents tau from binding to microtubules and promotes microtubule disassembly. Finally, we show that heparan sulfate and hyperphosphorylated tau coexist in nerve cells of the Alzheimer's disease brain at the earliest known stages of neurofibrillary pathol. These findings, with previous studies which show that heparin stimulates tau phosphorylation by a no. of protein kinases, indicate that sulfated glycosaminoglycans may be a key factor in the formation of the neurofibrillary lesions of Alzheimer's disease.
- 6Fichou, Y.; Lin, Y.; Rauch, J. N.; Vigers, M.; Zeng, Z.; Srivastava, M.; Keller, T. J.; Freed, J. H.; Kosik, K. S.; Han, S. Cofactors are essential constituents of stable and seeding-active tau fibrils. Proc. Natl. Acad. Sci. U.S.A. 2018, 115, 13234– 13239, DOI: 10.1073/pnas.1810058115Google Scholar6https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXisF2rsrvL&md5=47c4d658171711202a8be7cbddf4535fCofactors are essential constituents of stable and seeding-active tau fibrilsFichou, Yann; Lin, Yanxian; Rauch, Jennifer N.; Vigers, Michael; Zeng, Zhikai; Srivastava, Madhur; Keller, Timothy J.; Freed, Jack H.; Kosik, Kenneth S.; Han, SongiProceedings of the National Academy of Sciences of the United States of America (2018), 115 (52), 13234-13239CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)Amyloid fibrils are cross-β-rich aggregates that are exceptionally stable forms of protein assembly. Accumulation of tau amyloid fibrils is involved in many neurodegenerative diseases, including Alzheimer's disease (AD). Heparin-induced aggregates have been widely used and assumed to be a good tau amyloid fibril model for most biophys. studies. Here we show that mature fibrils made of 4R tau variants, prepd. with heparin or RNA, spontaneously depolymerize and release monomers when their cofactors are removed. We demonstrate that the cross-β-sheet assembly formed in vitro with polyanion addn. is unstable at room temp. We furthermore demonstrate high seeding capacity with transgenic AD mouse brain-extd. tau fibrils in vitro that, however, is exhausted after one generation, while supplementation with RNA cofactors resulted in sustained seeding over multiple generations. We suggest that tau fibrils formed in brains are supported by unknown cofactors and inhere higher-quality packing, as reflected in a more distinct conformational arrangement in the mouse fibril-seeded, compared with heparin-induced, tau fibrils. Our study suggests that the role of cofactors in tauopathies is a worthy focus of future studies, as they may be viable targets for diagnosis and therapeutics.
- 7Achazi, K.; Haag, R.; Ballauff, M.; Dernedde, J.; Kizhakkedathu, J. N.; Maysinger, D.; Multhaup, G. Understanding the interaction of polyelectrolyte architectures with proteins and biosystems. Angew. Chem., Int. Ed. 2021, 60, 3882– 3904, DOI: 10.1002/anie.202006457Google Scholar7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXitFyhs7nM&md5=84aa4dbce05059373ed72fc46bd469cbUnderstanding the Interaction of Polyelectrolyte Architectures with Proteins and BiosystemsAchazi, Katharina; Haag, Rainer; Ballauff, Matthias; Dernedde, Jens; Kizhakkedathu, Jayachandran N.; Maysinger, Dusica; Multhaup, GerdAngewandte Chemie, International Edition (2021), 60 (8), 3882-3904CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. The counterions neutralizing the charges on polyelectrolytes such as DNA or heparin may dissoc. in water and greatly influence the interaction of such polyelectrolytes with biomols., particularly proteins. In this Review the authors give an overview of studies on the interaction of proteins with polyelectrolytes and how this knowledge can be used for medical applications. Counterion release was identified as the main driving force for the binding of proteins to polyelectrolytes: Patches of pos. charge become multivalent counterions of the polyelectrolyte and lead to the release of counterions from the polyelectrolyte and a concomitant increase in entropy. This is shown from studies on the interaction of proteins with natural and synthetic polyelectrolytes. Special emphasis is paid to sulfated dendritic polyglycerols (dPGS). The Review demonstrates that the authors are moving to a better understanding of charge-charge interactions in systems of biol. relevance. Research along these lines will aid and promote the design of synthetic polyelectrolytes for medical applications.
- 8Kayitmazer, A. B.; Seeman, D.; Minsky, B. B.; Dubin, P. L.; Xu, Y. Protein–polyelectrolyte interactions. Soft Matter 2013, 9, 2553– 2583, DOI: 10.1039/c2sm27002aGoogle Scholar8https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXitVCitr0%253D&md5=652b0fe09debf23e810f733326abab58Protein-polyelectrolyte interactionsKayitmazer, A. Basak; Seeman, Daniel; Minsky, Burcu Baykal; Dubin, Paul L.; Xu, YishengSoft Matter (2013), 9 (9), 2553-2583CODEN: SMOABF; ISSN:1744-683X. (Royal Society of Chemistry)A review. The interactions of proteins and polyelectrolytes lead to diverse applications in sepns., delivery and wound repair, and are thus of interest to scientists in e.g. (a) glycobiol., (b) tissue engineering, (c) biosensing, and (d) pharmacol. This breadth is accompanied by an assortment of contexts and models in which polyelectrolytes are seen as (a) protein cognates assisting in complex cellular roles, (b) surrogates for the extracellular matrix, mimicking its hydration, mech. and sequestering properties, (c) benign hosts that gently entrap, deposit and tether protein substrate specificity, and (d) selective but non-specific agents that modify protein soly. Unsurprisingly, this literature is somewhat segregated by objectives and paradigms. We hope this review, which emphasizes publications over the last 8 years, represents and also counterbalances that divergence. An ongoing theme is the role of electrostatics, and we show how this leads to the variety of phys. forms taken by protein-polyelectrolyte complexes. We present approaches towards anal. and characterization, motivated by the goal of structure-property elucidation. Such understanding should guide in applications, our third topic. We present recent developments in modeling and simulations of protein-polyelectrolyte systems. We close with a prospective on future developments in this field.
- 9Nandy, B.; Saurabh, S.; Sahoo, A. K.; Dixit, N. M.; Maiti, P. K. The SPL7013 dendrimer destabilizes the HIV-1 gp120-CD4 complex. Nanoscale 2015, 7, 18628– 18641, DOI: 10.1039/C5NR04632GGoogle Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhs1Squ7bF&md5=3a1bef77f1f9ac6af6f04778be48045eThe SPL7013 dendrimer destabilizes the HIV-1 gp120-CD4 complexNandy, Bidisha; Saurabh, Suman; Sahoo, Anil Kumar; Dixit, Narendra M.; Maiti, Prabal K.Nanoscale (2015), 7 (44), 18628-18641CODEN: NANOHL; ISSN:2040-3372. (Royal Society of Chemistry)The poly (l-lysine)-based SPL7013 dendrimer with naphthalene disulfonate surface groups blocks the entry of HIV-1 into target cells and is in clin. trials for development as a topical microbicide. Its mechanism of action against R5 HIV-1, the HIV-1 variant implicated in transmission across individuals, remains poorly understood. Using docking and fully atomistic MD simulations, we find that SPL7013 binds tightly to R5 gp120 in the gp120-CD4 complex but weakly to gp120 alone. Further, the binding, although to multiple regions of gp120, does not occlude the CD4 binding site on gp120, suggesting that SPL7013 does not prevent the binding of R5 gp120 to CD4. Using MD simulations to compute binding energies of several docked structures, we find that SPL7013 binding to gp120 significantly weakens the gp120-CD4 complex. Finally, we use steered mol. dynamics (SMD) to study the kinetics of the dissocn. of the gp120-CD4 complex in the absence of the dendrimer and with the dendrimer bound in each of the several stable configurations to gp120. We find that SPL7013 significantly lowers the force required to rupture the gp120-CD4 complex and accelerates its dissocn. Taken together, our findings suggest that SPL7013 compromises the stability of the R5 gp120-CD4 complex, potentially preventing the accrual of the requisite no. of gp120-CD4 complexes across the virus-cell interface, thereby blocking virus entry.
- 10Cagno, V.; Tseligka, E. D.; Jones, S. T.; Tapparel, C. Heparan sulfate proteoglycans and viral attachment: true receptors or adaptation bias?. Viruses 2019, 11, 596 DOI: 10.3390/v11070596Google Scholar10https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXisVOgu7jO&md5=3cf85bf40431c53df1a530498ddabf4aHeparan sulfate proteoglycans and viral attachment: true receptors or adaptation bias?Cagno, Valeria; Tseligka, Eirini D.; Jones, Samuel T.; Tapparel, CarolineViruses (2019), 11 (7), 596CODEN: VIRUBR; ISSN:1999-4915. (MDPI AG)A review. Heparan sulfate proteoglycans (HSPG) are composed of unbranched, neg. charged heparan sulfate (HS) polysaccharides attached to a variety of cell surface or extracellular matrix proteins. Widely expressed, they mediate many biol. activities, including angiogenesis, blood coagulation, developmental processes, and cell homeostasis. HSPG are highly sulfated and broadly used by a range of pathogens, esp. viruses, to attach to the cell surface. In this review, we summarize the current knowledge on HSPG-virus interactions and distinguish viruses with established HS binding, viruses that bind HS only after intra-host or cell culture adaptation, and finally, viruses whose dependence on HS for infection is debated. We also provide an overview of the antiviral compds. designed to interfere with HS binding. Many questions remain about the true importance of these receptors in vivo, knowledge that is crit. for the design of future antiviral therapies.
- 11Liese, S.; Netz, R. R. Quantitative prediction of multivalent ligand-receptor binding affinities for influenza, cholera, and anthrax inhibition. ACS Nano 2018, 12, 4140– 4147, DOI: 10.1021/acsnano.7b08479Google Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXjtlOnsr8%253D&md5=8a1df12e15cf156042bec1a0cadc51daQuantitative Prediction of Multivalent Ligand-Receptor Binding Affinities for Influenza, Cholera, and Anthrax InhibitionLiese, Susanne; Netz, Roland R.ACS Nano (2018), 12 (5), 4140-4147CODEN: ANCAC3; ISSN:1936-0851. (American Chemical Society)Multivalency achieves strong, yet reversible binding by the simultaneous formation of multiple weak bonds. It is a key interaction principle in biol. and promising for the synthesis of high-affinity inhibitors of pathogens. We present a mol. model for the binding affinity of synthetic multivalent ligands onto multivalent receptors consisting of n receptor units arranged on a regular polygon. Ligands consist of a geometrically matching rigid polygonal core to which monovalent ligand units are attached via flexible linker polymers, closely mimicking existing exptl. designs. The calcd. binding affinities quant. agree with exptl. studies for cholera toxin (n = 5) and anthrax receptor (n = 7) and allow to predict optimal core size and optimal linker length. Maximal binding affinity is achieved for a core that matches the receptor size and for linkers that have an equil. end-to-end distance that is slightly longer than the geometric sepn. between ligand core and receptor sites. Linkers that are longer than optimal are greatly preferable compared to shorter linkers. The angular steric restriction between ligand unit and linker polymer is shown to be a key parameter. We construct an enhancement diagram that quantifies the multivalent binding affinity compared to monovalent ligands. We conclude that multivalent ligands against influenza viral hemagglutinin (n = 3), cholera toxin (n = 5), and anthrax receptor (n = 7) can outperform monovalent ligands only for a monovalent ligand affinity that exceeds a core-size dependent threshold value. Thus, multivalent drug design needs to balance core size, linker length, as well as monovalent ligand unit affinity.
- 12Lauster, D.; Osterrieder, K.; Haag, R.; Ballauff, M.; Herrmann, A. Respiratory viruses interacting with cells: The importance of electrostatics. Front. Microbiol. 2023, 14, 1169547 DOI: 10.3389/fmicb.2023.1169547Google ScholarThere is no corresponding record for this reference.
- 13Chu, H.; Hu, B.; Huang, X.; Chai, Y.; Zhou, D.; Wang, Y.; Shuai, H.; Yang, D.; Hou, Y.; Zhang, X. Host and viral determinants for efficient SARS-CoV-2 infection of the human lung. Nat. Commun. 2021, 12, 134 DOI: 10.1038/s41467-020-20457-wGoogle Scholar13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhtF2gurg%253D&md5=637c93ff402b13e3a45c2e5dc53548d7Host and viral determinants for efficient SARS-CoV-2 infection of the human lungChu, Hin; Hu, Bingjie; Huang, Xiner; Chai, Yue; Zhou, Dongyan; Wang, Yixin; Shuai, Huiping; Yang, Dong; Hou, Yuxin; Zhang, Xi; Yuen, Terrence Tsz-Tai; Cai, Jian-Piao; Zhang, Anna Jinxia; Zhou, Jie; Yuan, Shuofeng; To, Kelvin Kai-Wang; Chan, Ivy Hau-Yee; Sit, Ko-Yung; Foo, Dominic Chi-Chung; Wong, Ian Yu-Hong; Ng, Ada Tsui-Lin; Cheung, Tan To; Law, Simon Ying-Kit; Au, Wing-Kuk; Brindley, Melinda A.; Chen, Zhiwei; Kok, Kin-Hang; Chan, Jasper Fuk-Woo; Yuen, Kwok-YungNature Communications (2021), 12 (1), 134CODEN: NCAOBW; ISSN:2041-1723. (Nature Research)Abstr.: Understanding the factors that contribute to efficient SARS-CoV-2 infection of human cells may provide insights on SARS-CoV-2 transmissibility and pathogenesis, and reveal targets of intervention. Here, we analyze host and viral determinants essential for efficient SARS-CoV-2 infection in both human lung epithelial cells and ex vivo human lung tissues. We identify heparan sulfate as an important attachment factor for SARS-CoV-2 infection. Next, we show that sialic acids present on ACE2 prevent efficient spike/ACE2-interaction. While SARS-CoV infection is substantially limited by the sialic acid-mediated restriction in both human lung epithelial cells and ex vivo human lung tissues, infection by SARS-CoV-2 is limited to a lesser extent. We further demonstrate that the furin-like cleavage site in SARS-CoV-2 spike is required for efficient virus replication in human lung but not intestinal tissues. These findings provide insights on the efficient SARS-CoV-2 infection of human lungs.
- 14Kim, S. H.; Kearns, F. L.; Rosenfeld, M. A.; Votapka, L.; Casalino, L.; Papanikolas, M.; Amaro, R. E.; Freeman, R. SARS-CoV-2 evolved variants optimize binding to cellular glycocalyx. Cell Rep. Phys. Sci. 2023, 4, 101346 DOI: 10.1016/j.xcrp.2023.101346Google ScholarThere is no corresponding record for this reference.
- 15Nie, C.; Pouyan, P.; Lauster, D.; Trimpert, J.; Kerkhoff, Y.; Szekeres, G. P.; Wallert, M.; Block, S.; Sahoo, A. K.; Dernedde, J.; Pagel, K.; Kaufer, B. B.; Netz, R. R.; Ballauff, M.; Haag, R. Polysulfates block SARS-CoV-2 uptake via electrostatic interactions. Angew. Chem., Int. Ed. 2021, 60, 15870– 15878, DOI: 10.1002/anie.202102717Google Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhsVKmtb7I&md5=681dca8bfea352c705f4ece7f110c6e3Polysulfates Block SARS-CoV-2 Uptake through Electrostatic InteractionsNie, Chuanxiong; Pouyan, Paria; Lauster, Daniel; Trimpert, Jakob; Kerkhoff, Yannic; Szekeres, Gergo Peter; Wallert, Matthias; Block, Stephan; Sahoo, Anil Kumar; Dernedde, Jens; Pagel, Kevin; Kaufer, Benedikt B.; Netz, Roland R.; Ballauff, Matthias; Haag, RainerAngewandte Chemie, International Edition (2021), 60 (29), 15870-15878CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)Here we report that neg. charged polysulfates can bind to the spike protein of SARS-CoV-2 via electrostatic interactions. Using a plaque redn. assay, we compare inhibition of SARS-CoV-2 by heparin, pentosan sulfate, linear polyglycerol sulfate (LPGS) and hyperbranched polyglycerol sulfate (HPGS). Highly sulfated LPGS is the optimal inhibitor, with an IC50 of 67μg mL-1 (approx. 1.6μM). This synthetic polysulfate exhibits more than 60-fold higher virus inhibitory activity than heparin (IC50: 4084μg mL-1), along with much lower anticoagulant activity. Furthermore, in mol. dynamics simulations, we verified that LPGS can bind more strongly to the spike protein than heparin, and that LPGS can interact even more with the spike protein of the new N501Y and E484K variants. Our study demonstrates that the entry of SARS-CoV-2 into host cells can be blocked via electrostatic interactions, therefore LPGS can serve as a blueprint for the design of novel viral inhibitors of SARS-CoV-2.
- 16Clausen, T. M.; Sandoval, D. R.; Spliid, C. B.; Pihl, J.; Perrett, H. R.; Painter, C. D.; Narayanan, A.; Majowicz, S. A.; Kwong, E. M.; McVicar, R. N. SARS-CoV-2 infection depends on cellular heparan sulfate and ACE2. Cell 2020, 183, 1043– 1057, DOI: 10.1016/j.cell.2020.09.033Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhvFGqsLjK&md5=f478b8b320b347ca7f8accca25b822bcSARS-CoV-2 Infection Depends on Cellular Heparan Sulfate and ACE2Clausen, Thomas Mandel; Sandoval, Daniel R.; Spliid, Charlotte B.; Pihl, Jessica; Perrett, Hailee R.; Painter, Chelsea D.; Narayanan, Anoop; Majowicz, Sydney A.; Kwong, Elizabeth M.; McVicar, Rachael N.; Thacker, Bryan E.; Glass, Charles A.; Yang, Zhang; Torres, Jonathan L.; Golden, Gregory J.; Bartels, Phillip L.; Porell, Ryan N.; Garretson, Aaron F.; Laubach, Logan; Feldman, Jared; Yin, Xin; Pu, Yuan; Hauser, Blake M.; Caradonna, Timothy M.; Kellman, Benjamin P.; Martino, Cameron; Gordts, Philip L. S. M.; Chanda, Sumit K.; Schmidt, Aaron G.; Godula, Kamil; Leibel, Sandra L.; Jose, Joyce; Corbett, Kevin D.; Ward, Andrew B.; Carlin, Aaron F.; Esko, Jeffrey D.Cell (Cambridge, MA, United States) (2020), 183 (4), 1043-1057.e15CODEN: CELLB5; ISSN:0092-8674. (Cell Press)We show that SARS-CoV-2 spike protein interacts with both cellular heparan sulfate and angiotensin-converting enzyme 2 (ACE2) through its receptor-binding domain (RBD). Docking studies suggest a heparin/heparan sulfate-binding site adjacent to the ACE2-binding site. Both ACE2 and heparin can bind independently to spike protein in vitro, and a ternary complex can be generated using heparin as a scaffold. Electron micrographs of spike protein suggests that heparin enhances the open conformation of the RBD that binds ACE2. On cells, spike protein binding depends on both heparan sulfate and ACE2. Unfractionated heparin, non-anticoagulant heparin, heparin lyases, and lung heparan sulfate potently block spike protein binding and/or infection by pseudotyped virus and authentic SARS-CoV-2 virus. We suggest a model in which viral attachment and infection involves heparan sulfate-dependent enhancement of binding to ACE2. Manipulation of heparan sulfate or inhibition of viral adhesion by exogenous heparin presents new therapeutic opportunities.
- 17Lever, R.; Page, C. P. Non-Anticoagulant Effects of Heparin: An Overview. In Heparin - A Century of Progress; Lever, R.; Mulloy, B.; Page, C. P., Eds.; Springer: Berlin, Heidelberg, 2012; pp 281– 305.Google ScholarThere is no corresponding record for this reference.
- 18Page, C. Heparin and related drugs: beyond anticoagulant activity. ISRN Pharmacol. 2013, 1, 910743 DOI: 10.1155/2013/910743Google ScholarThere is no corresponding record for this reference.
- 19Oates, J. A.; Wood, A. J. J.; Hirsh, J. Heparin. N. Engl. J. Med. 1991, 324, 1565– 1574, DOI: 10.1056/NEJM199105303242206Google ScholarThere is no corresponding record for this reference.
- 20Cate, H. T. Surviving Covid-19 with Heparin?. N. Engl. J. Med. 2021, 385, 845– 846, DOI: 10.1056/NEJMe2111151Google ScholarThere is no corresponding record for this reference.
- 21Lan, J.; Ge, J.; Yu, J.; Shan, S.; Zhou, H.; Fan, S.; Zhang, Q.; Shi, X.; Wang, Q.; Zhang, L.; Wang, X. Structure of the SARS-CoV-2 spike receptor-binding domain bound to the ACE2 receptor. Nature 2020, 581, 215– 220, DOI: 10.1038/s41586-020-2180-5Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXoslOqtL8%253D&md5=279c60143e8e5eb505457e0778baa8efStructure of the SARS-CoV-2 spike receptor-binding domain bound to the ACE2 receptorLan, Jun; Ge, Jiwan; Yu, Jinfang; Shan, Sisi; Zhou, Huan; Fan, Shilong; Zhang, Qi; Shi, Xuanling; Wang, Qisheng; Zhang, Linqi; Wang, XinquanNature (London, United Kingdom) (2020), 581 (7807), 215-220CODEN: NATUAS; ISSN:0028-0836. (Nature Research)Abstr.: A new and highly pathogenic coronavirus (severe acute respiratory syndrome coronavirus-2, SARS-CoV-2) caused an outbreak in Wuhan city, Hubei province, China, starting from Dec. 2019 that quickly spread nationwide and to other countries around the world1-3. Here, to better understand the initial step of infection at an at. level, we detd. the crystal structure of the receptor-binding domain (RBD) of the spike protein of SARS-CoV-2 bound to the cell receptor ACE2. The overall ACE2-binding mode of the SARS-CoV-2 RBD is nearly identical to that of the SARS-CoV RBD, which also uses ACE2 as the cell receptor4. Structural anal. identified residues in the SARS-CoV-2 RBD that are essential for ACE2 binding, the majority of which either are highly conserved or share similar side chain properties with those in the SARS-CoV RBD. Such similarity in structure and sequence strongly indicate convergent evolution between the SARS-CoV-2 and SARS-CoV RBDs for improved binding to ACE2, although SARS-CoV-2 does not cluster within SARS and SARS-related coronaviruses1-3,5. The epitopes of two SARS-CoV antibodies that target the RBD are also analyzed for binding to the SARS-CoV-2 RBD, providing insights into the future identification of cross-reactive antibodies.
- 22Pishko, A. M.; Lefler, D. S.; Gimotty, P.; Paydary, K.; Fardin, S.; Arepally, G. M.; Crowther, M.; Rice, L.; Vega, R.; Cines, D. B. The risk of major bleeding in patients with suspected heparin-induced thrombocytopenia. J. Thromb. Haemostasis 2019, 17, 1956– 1965, DOI: 10.1111/jth.14587Google ScholarThere is no corresponding record for this reference.
- 23Nie, C.; Sahoo, A. K.; Netz, R. R.; Herrmann, A.; Ballauff, M.; Haag, R. Charge matters: Mutations in omicron variant favor binding to cells. ChemBioChem 2022, 23, e202100681 DOI: 10.1002/cbic.202100681Google Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB38Xislenu7w%253D&md5=69b8f1e9b595a12dbb6ed73d09445a2dCharge Matters: Mutations in Omicron Variant Favor Binding to CellsNie, Chuanxiong; Sahoo, Anil Kumar; Netz, Roland R.; Herrmann, Andreas; Ballauff, Matthias; Haag, RainerChemBioChem (2022), 23 (6), e202100681CODEN: CBCHFX; ISSN:1439-4227. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Evidence is strengthening to suggest that the novel SARS-CoV-2 mutant Omicron, with its more than 60 mutations, will spread and dominate worldwide. Although the mutations in the spike protein are known, the mol. basis for why the addnl. mutations in the spike protein that have not previously occurred account for Omicron's higher infection potential, is not understood. We propose, based on chem. rational and mol. dynamics simulations, that the elevated occurrence of pos. charged amino acids in certain domains of the spike protein (Delta: +4; Omicron: +5 vs. wild type) increases binding to cellular polyanionic receptors, such as heparan sulfate due to multivalent charge-charge interactions. This observation is a starting point for targeted drug development.
- 24Wang, L.; Wu, Y.; Deng, Y.; Kim, B.; Pierce, L.; Krilov, G.; Lupyan, D.; Robinson, S.; Dahlgren, M. K.; Greenwood, J. Accurate and reliable prediction of relative ligand binding potency in prospective drug discovery by way of a modern free-energy calculation protocol and force field. J. Am. Chem. Soc. 2015, 137, 2695– 2703, DOI: 10.1021/ja512751qGoogle Scholar24https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhsF2iuro%253D&md5=37a4f4a6c085f47ed531342643b6c33bAccurate and Reliable Prediction of Relative Ligand Binding Potency in Prospective Drug Discovery by Way of a Modern Free-Energy Calculation Protocol and Force FieldWang, Lingle; Wu, Yujie; Deng, Yuqing; Kim, Byungchan; Pierce, Levi; Krilov, Goran; Lupyan, Dmitry; Robinson, Shaughnessy; Dahlgren, Markus K.; Greenwood, Jeremy; Romero, Donna L.; Masse, Craig; Knight, Jennifer L.; Steinbrecher, Thomas; Beuming, Thijs; Damm, Wolfgang; Harder, Ed; Sherman, Woody; Brewer, Mark; Wester, Ron; Murcko, Mark; Frye, Leah; Farid, Ramy; Lin, Teng; Mobley, David L.; Jorgensen, William L.; Berne, Bruce J.; Friesner, Richard A.; Abel, RobertJournal of the American Chemical Society (2015), 137 (7), 2695-2703CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Designing tight-binding ligands is a primary objective of small-mol. drug discovery. Over the past few decades, free-energy calcns. have benefited from improved force fields and sampling algorithms, as well as the advent of low-cost parallel computing. However, it has proven to be challenging to reliably achieve the level of accuracy that would be needed to guide lead optimization (∼5× in binding affinity) for a wide range of ligands and protein targets. Not surprisingly, widespread com. application of free-energy simulations has been limited due to the lack of large-scale validation coupled with the tech. challenges traditionally assocd. with running these types of calcns. Here, we report an approach that achieves an unprecedented level of accuracy across a broad range of target classes and ligands, with retrospective results encompassing 200 ligands and a wide variety of chem. perturbations, many of which involve significant changes in ligand chem. structures. In addn., we have applied the method in prospective drug discovery projects and found a significant improvement in the quality of the compds. synthesized that have been predicted to be potent. Compds. predicted to be potent by this approach have a substantial redn. in false positives relative to compds. synthesized on the basis of other computational or medicinal chem. approaches. Furthermore, the results are consistent with those obtained from our retrospective studies, demonstrating the robustness and broad range of applicability of this approach, which can be used to drive decisions in lead optimization.
- 25Cournia, Z.; Allen, B.; Sherman, W. Relative binding free energy calculations in drug discovery: recent advances and practical considerations. J. Chem. Inf. Model. 2017, 57, 2911– 2937, DOI: 10.1021/acs.jcim.7b00564Google Scholar25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhvFOhu7fM&md5=ffcc9f970ab660eced8d3343ccefeec6Relative Binding Free Energy Calculations in Drug Discovery: Recent Advances and Practical ConsiderationsCournia, Zoe; Allen, Bryce; Sherman, WoodyJournal of Chemical Information and Modeling (2017), 57 (12), 2911-2937CODEN: JCISD8; ISSN:1549-9596. (American Chemical Society)A review. Accurate in silico prediction of protein-ligand binding affinities has been a primary objective of structure-based drug design for decades due to the putative value it would bring to the drug discovery process. However, computational methods have historically failed to deliver value in real-world drug discovery applications due to a variety of scientific, tech., and practical challenges. Recently, a family of approaches commonly referred to as relative binding free energy (RBFE) calcns., which rely on physics-based mol. simulations and statistical mechanics, have shown promise in reliably generating accurate predictions in the context of drug discovery projects. This advance arises from accumulating developments in the underlying scientific methods (decades of research on force fields and sampling algorithms) coupled with vast increases in computational resources (graphics processing units and cloud infrastructures). Mounting evidence from retrospective validation studies, blind challenge predictions, and prospective applications suggests that RBFE simulations can now predict the affinity differences for congeneric ligands with sufficient accuracy and throughput to deliver considerable value in hit-to-lead and lead optimization efforts. Here, the authors present an overview of current RBFE implementations, highlighting recent advances and remaining challenges, along with examples that emphasize practical considerations for obtaining reliable RBFE results. The authors focus specifically on relative binding free energies because the calcns. are less computationally intensive than abs. binding free energy (ABFE) calcns. and map directly onto the hit-to-lead and lead optimization processes, where the prediction of relative binding energies between a ref. mol. and new ideas (virtual mols.) can be used to prioritize mols. for synthesis. The authors describe the crit. aspects of running RBFE calcns., from both theor. and applied perspectives, using a combination of retrospective literature examples and prospective studies from drug discovery projects. This work is intended to provide a contemporary overview of the scientific, tech., and practical issues assocd. with running relative binding free energy simulations, with a focus on real-world drug discovery applications. The authors offer guidelines for improving the accuracy of RBFE simulations, esp. for challenging cases, and emphasize unresolved issues that could be improved by further research in the field.
- 26Mobley, D. L.; Gilson, M. K. Predicting binding free energies: frontiers and benchmarks. Annu. Rev. Biophys. 2017, 46, 531– 558, DOI: 10.1146/annurev-biophys-070816-033654Google Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXlvFClsr0%253D&md5=38af2949ba205daaf9899b87c4e1a2bdPredicting Binding Free Energies: Frontiers and BenchmarksMobley, David L.; Gilson, Michael K.Annual Review of Biophysics (2017), 46 (), 531-558CODEN: ARBNCV; ISSN:1936-122X. (Annual Reviews)Binding free energy calcns. based on mol. simulations provide predicted affinities for biomol. complexes. These calcns. begin with a detailed description of a system, including its chem. compn. and the interactions among its components. Simulations of the system are then used to compute thermodn. information, such as binding affinities. Because of their promise for guiding mol. design, these calcns. have recently begun to see widespread applications in early-stage drug discovery. However, many hurdles remain in making them a robust and reliable tool. In this review, we highlight key challenges of these calcns., discuss some examples of these challenges, and call for the designation of std. community benchmark test systems that will help the research community generate and evaluate progress. In our view, progress will require careful assessment and evaluation of new methods, force fields, and modeling innovations on well-characterized benchmark systems, and we lay out our vision for how this can be achieved.
- 27Zhao, J.; Cao, Y.; Zhang, L. Exploring the computational methods for protein-ligand binding site prediction. Comput. Struct. Biotechnol. J. 2020, 18, 417– 426, DOI: 10.1016/j.csbj.2020.02.008Google Scholar27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXktFeitrY%253D&md5=099de1ae131781598f5d09d48dd23061Exploring the computational methods for protein-ligand binding site predictionZhao, Jingtian; Cao, Yang; Zhang, LeComputational and Structural Biotechnology Journal (2020), 18 (), 417-426CODEN: CSBJAC; ISSN:2001-0370. (Elsevier B.V.)A review. Proteins participate in various essential processes in vivo via interactions with other mols. Identifying the residues participating in these interactions not only provides biol. insights for protein function studies but also has great significance for drug discoveries. Therefore, predicting protein-ligand binding sites has long been under intense research in the fields of bioinformatics and computer aided drug discovery. In this review, we first introduce the research background of predicting protein-ligand binding sites and then classify the methods into four categories, namely, 3D structure-based, template similarity-based, traditional machine learning-based and deep learning-based methods. We describe representative algorithms in each category and elaborate on machine learning and deep learning-based prediction methods in more detail. Finally, we discuss the trends and challenges of the current research such as mol. dynamics simulation based cryptic binding sites prediction, and highlight prospective directions for the near future.
- 28Gapsys, V.; Yildirim, A.; Aldeghi, M.; Khalak, Y.; Van der Spoel, D.; de Groot, B. L. Accurate absolute free energies for ligand–protein binding based on non-equilibrium approaches. Commun. Chem. 2021, 4, 61 DOI: 10.1038/s42004-021-00498-yGoogle Scholar28https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhvFSktbzO&md5=212f9b8a8ca965e71ec680daa550183fAccurate absolute free energies for ligand-protein binding based on non-equilibrium approachesGapsys, Vytautas; Yildirim, Ahmet; Aldeghi, Matteo; Khalak, Yuriy; van der Spoel, David; de Groot, Bert L.Communications Chemistry (2021), 4 (1), 61CODEN: CCOHCT; ISSN:2399-3669. (Nature Research)Abstr.: The accurate calcn. of the binding free energy for arbitrary ligand-protein pairs is a considerable challenge in computer-aided drug discovery. Recently, it has been demonstrated that current state-of-the-art mol. dynamics (MD) based methods are capable of making highly accurate predictions. Conventional MD-based approaches rely on the first principles of statistical mechanics and assume equil. sampling of the phase space. In the current work we demonstrate that accurate abs. binding free energies (ABFE) can also be obtained via theor. rigorous non-equil. approaches. Our investigation of ligands binding to bromodomains and T4 lysozyme reveals that both equil. and non-equil. approaches converge to the same results. The non-equil. approach achieves the same level of accuracy and convergence as an equil. free energy perturbation (FEP) method enhanced by Hamiltonian replica exchange. We also compare uni- and bi-directional non-equil. approaches and demonstrate that considering the work distributions from both forward and reverse directions provides substantial accuracy gains. In summary, non-equil. ABFE calcns. are shown to yield reliable and well-converged ests. of protein-ligand binding affinity.
- 29Xu, X.; Angioletti-Uberti, S.; Lu, Y.; Dzubiella, J.; Ballauff, M. Interaction of proteins with polyelectrolytes: Comparison of theory to experiment. Langmuir 2019, 35, 5373– 5391, DOI: 10.1021/acs.langmuir.8b01802Google Scholar29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhsVOltrvM&md5=843d9f0f903cbc530b2dcc43ff6118a6Interaction of Proteins with Polyelectrolytes: Comparison of Theory to ExperimentXu, Xiao; Angioletti-Uberti, Stefano; Lu, Yan; Dzubiella, Joachim; Ballauff, MatthiasLangmuir (2019), 35 (16), 5373-5391CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)A review on the authors' recent studies on the interaction of simple proteins such as human serum albumin (HSA) and lysozyme with linear polyelectrolytes, charged dendrimers, charged networks, and polyelectrolyte brushes, based on exptl. works combined with mol. dynamics (MD) simulations and mean-field theories.
- 30Paiardi, G.; Ferraz, M.; Rusnati, M.; Wade, R. C. The accomplices: Heparan sulfates and N-glycans foster SARS-CoV-2 spike: ACE2 receptor binding and virus priming. Proc. Natl. Acad. Sci. U.S.A. 2024, 121, e2404892121 DOI: 10.1073/pnas.2404892121Google ScholarThere is no corresponding record for this reference.
- 31Bhatia, S.; Lauster, D.; Bardua, M.; Ludwig, K.; Angioletti-Uberti, S.; Popp, N.; Hoffmann, U.; Paulus, F.; Budt, M.; Stadtmüller, M. Linear polysialoside outperforms dendritic analogs for inhibition of influenza virus infection in vitro and in vivo. Biomaterials 2017, 138, 22– 34, DOI: 10.1016/j.biomaterials.2017.05.028Google Scholar31https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXosVWnsL8%253D&md5=409628a080116660d39f1a1a41cd6c7bLinear polysialoside outperforms dendritic analogs for inhibition of influenza virus infection in vitro and in vivoBhatia, Sumati; Lauster, Daniel; Bardua, Markus; Ludwig, Kai; Angioletti-Uberti, Stefano; Popp, Nicole; Hoffmann, Ute; Paulus, Florian; Budt, Matthias; Stadtmueller, Marlena; Wolff, Thorsten; Hamann, Alf; Boettcher, Christoph; Herrmann, Andreas; Haag, RainerBiomaterials (2017), 138 (), 22-34CODEN: BIMADU; ISSN:0142-9612. (Elsevier Ltd.)Inhibition of influenza A virus infection by multivalent sialic acid inhibitors preventing viral hemagglutinin binding to host cells of the respiratory tract is a promising strategy. However, optimal geometry and optimal ligand presentation on multivalent scaffolds for efficient inhibition both in vitro and in vivo application are still unclear. Here, by comparing linear and dendritic polyglycerol sialosides (LPGSA and dPGSA) we identified architectural requirements and optimal ligand densities for an efficient multivalent inhibitor of influenza virus A/X31/1 (H3N2). Due to its large vol., the LPGSA at optimal ligand d. sterically shielded the virus significantly better than the dendritic analog. A statistical mechanics model rationalizes the relevance of ligand d., morphol., and the size of multivalent scaffolds for the potential to inhibit virus-cell binding. Optimized LPGSA inhibited virus infection at IC50 in the low nanomolar nanoparticle concn. range and also showed potent antiviral activity against two avian influenza strains A/Mallard/439/2004 (H3N2) and A/turkey/Italy/472/1999 (H7N1) post infection. In vivo application of inhibitors clearly confirmed the higher inhibition potential of linear multivalent scaffolds to prevent infection. The optimized LPGSA did not show any acute toxicity, and was much more potent than the neuraminidase inhibitor oseltamivir carboxylate in vivo. Combined application of the LPGSA and oseltamivir carboxylate revealed a synergistic inhibitory effect and successfully prevented influenza virus infection in mice.
- 32Xu, C.; Wang, Y.; Liu, C.; Zhang, C.; Han, W.; Hong, X.; Wang, Y.; Hong, Q.; Wang, S.; Zhao, Q. Conformational dynamics of SARS-CoV-2 trimeric spike glycoprotein in complex with receptor ACE2 revealed by cryo-EM. Sci. Adv. 2021, 7, eabe5575 DOI: 10.1126/sciadv.abe5575Google ScholarThere is no corresponding record for this reference.
- 33Horinek, D.; Serr, A.; Geisler, M.; Pirzer, T.; Slotta, U.; Lud, S. Q.; Garrido, J.; Scheibel, T.; Hugel, T.; Netz, R. R. Peptide adsorption on a hydrophobic surface results from an interplay of solvation, surface, and intrapeptide forces. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 2842– 2847, DOI: 10.1073/pnas.0707879105Google Scholar33https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXjtVSju70%253D&md5=8509147b175b299592be205938e8d48dPeptide adsorption on a hydrophobic surface results from an interplay of solvation, surface, and intrapeptide forcesHorinek, D.; Serr, A.; Geisler, M.; Pirzer, T.; Slotta, U.; Lud, S. Q.; Garrido, J. A.; Scheibel, T.; Hugel, T.; Netz, R. R.Proceedings of the National Academy of Sciences of the United States of America (2008), 105 (8), 2842-2847CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)The hydrophobic effect, i.e., the poor solvation of nonpolar parts of mols., plays a key role in protein folding and more generally for mol. self-assembly and aggregation in aq. media. The perturbation of the water structure accounts for many aspects of protein hydrophobicity. However, to what extent the dispersion interaction between mol. entities themselves contributes has remained unclear. This is so because in peptide folding interactions and structural changes occur on all length scales and make disentangling various contributions impossible. The authors address this issue both exptl. and theor. by looking at the force necessary to peel a mildly hydrophobic single peptide mol. from a flat hydrophobic diamond surface in the presence of water. This setup avoids problems caused by bubble adsorption, cavitation, and slow equilibration that complicate the much-studied geometry with two macroscopic surfaces. Using at.-force spectroscopy, the authors det. the mean desorption force of a single spider-silk peptide chain as F = 58 ± 8 pN, which corresponds to a desorption free energy of ≈5 k8T per amino acid. The authors' all-atomistic mol. dynamics simulation including explicit water correspondingly yields the desorption force F = 54 ± 15 pN. This observation demonstrates that std. nonpolarizable force fields used in classical simulations are capable of resolving the fine details of the hydrophobic attraction of peptides. The anal. of the involved energetics shows that water-structure effects and dispersive interactions give contributions of comparable magnitude that largely cancel out. It follows that the correct modeling of peptide hydrophobicity must take the intimate coupling of solvation and dispersive effects into account.
- 34Schwierz, N.; Horinek, D.; Liese, S.; Pirzer, T.; Balzer, B. N.; Hugel, T.; Netz, R. R. On the relationship between peptide adsorption resistance and surface contact angle: a combined experimental and simulation single-molecule study. J. Am. Chem. Soc. 2012, 134, 19628– 19638, DOI: 10.1021/ja304462uGoogle Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhsFygtb%252FJ&md5=ebb99ca67cd645f7d3553cf517a54ed3On the Relationship between Peptide Adsorption Resistance and Surface Contact Angle: A Combined Experimental and Simulation Single-Molecule StudySchwierz, Nadine; Horinek, Dominik; Liese, Susanne; Pirzer, Tobias; Balzer, Bizan N.; Hugel, Thorsten; Netz, Roland R.Journal of the American Chemical Society (2012), 134 (48), 19628-19638CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The force-induced desorption of single peptide chains from mixed OH/CH3-terminated self-assembled monolayers was studied in closely matched mol. dynamics simulations and at. force microscopy expts. with the goal to gain microscopic understanding of the transition between peptide adsorption and adsorption resistance as the surface contact angle is varied. In both simulations and expts., the surfaces become adsorption resistant against hydrophilic as well as hydrophobic peptides when their contact angle decreases below θ ≈ 50°-60°, thus confirming the so-called Berg limit established in the context of protein and cell adsorption. Entropy/enthalpy decompn. of the simulation results reveals that the key discriminator between the adsorption of different residues on a hydrophobic monolayer is of entropic nature and thus probably is linked to the hydrophobic effect. By pushing a polyalanine peptide onto a polar surface, simulations reveal that the peptide adsorption resistance is caused by the strongly bound water hydration layer and characterized by the simultaneous gain of both total entropy in the system and total no. of hydrogen bonds between water, peptide, and surface. This mechanistic insight into peptide adsorption resistance might help to refine design principles for anti-fouling surfaces.
- 35Page, T. M.; Nie, C.; Neander, L.; Povolotsky, T. L.; Sahoo, A. K.; Nickl, P.; Adler, J. M.; Bawadkji, O.; Radnik, J.; Achazi, K. Functionalized Fullerene for Inhibition of SARS-CoV-2 Variants. Small 2023, 19, 2206154 DOI: 10.1002/smll.202206154Google Scholar35https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3sXhsFylur0%253D&md5=ca14d876275ed6fe55ed8d23d417e1e6Functionalized Fullerene for Inhibition of SARS-CoV-2 VariantsPage, Taylor M.; Nie, Chuanxiong; Neander, Lenard; Povolotsky, Tatyana L.; Sahoo, Anil Kumar; Nickl, Philip; Adler, Julia M.; Bawadkji, Obida; Radnik, Jorg; Achazi, Katharina; Ludwig, Kai; Lauster, Daniel; Netz, Roland R.; Trimpert, Jakob; Kaufer, Benedikt; Haag, Rainer; Donskyi, Ievgen S.Small (2023), 19 (15), 2206154CODEN: SMALBC; ISSN:1613-6810. (Wiley-VCH Verlag GmbH & Co. KGaA)As virus outbreaks continue to pose a challenge, a nonspecific viral inhibitor can provide significant benefits, esp. against respiratory viruses. Polyglycerol sulfates recently emerge as promising agents that mediate interactions between cells and viruses through electrostatics, leading to virus inhibition. Similarly, hydrophobic C60 fullerene can prevent virus infection via interactions with hydrophobic cavities of surface proteins. Here, two strategies are combined to inhibit infection of SARS-CoV-2 variants in vitro. Effective inhibitory concns. in the millimolar range highlight the significance of bare fullerene's hydrophobic moiety and electrostatic interactions of polysulfates with surface proteins of SARS-CoV-2. Furthermore, microscale thermophoresis measurements support that fullerene linear polyglycerol sulfates interact with the SARS-CoV-2 virus via its spike protein, and highlight importance of electrostatic interactions within it. All-atom mol. dynamics simulations reveal that the fullerene binding site is situated close to the receptor binding domain, within 4 nm of polyglycerol sulfate binding sites, feasibly allowing both portions of the material to interact simultaneously.
- 36Woo, H.; Park, S.-J.; Choi, Y. K.; Park, T.; Tanveer, M.; Cao, Y.; Kern, N. R.; Lee, J.; Yeom, M. S.; Croll, T. I. Developing a fully glycosylated full-length SARS-CoV-2 spike protein model in a viral membrane. J. Phys. Chem. B 2020, 124, 7128– 7137, DOI: 10.1021/acs.jpcb.0c04553Google Scholar36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXht1WmsrjK&md5=a74433db7db78137b36d5fd714d49dd1Developing a fully glycosylated full-length SARS-CoV-2 spike protein model in a viral membraneWoo, Hyeonuk; Park, Sang-Jun; Choi, Yeol Kyo; Park, Taeyong; Tanveer, Maham; Cao, Yiwei; Kern, Nathan R.; Lee, Jumin; Yeom, Min Sun; Croll, Tristan I.; Seok, Chaok; Im, WonpilJournal of Physical Chemistry B (2020), 124 (33), 7128-7137CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)This tech. study describes all-atom modeling and simulation of a fully glycosylated full-length SARS-CoV-2 spike (S) protein in a viral membrane. First, starting from PDB: 6VSB and 6VXX, full-length S protein structures were modeled using template-based modeling, de-novo protein structure prediction, and loop modeling techniques in GALAXY modeling suite. Then, using the recently detd. most occupied glycoforms, 22 N-glycans and 1 O-glycan of each monomer were modeled using Glycan Reader & Modeler in CHARMM-GUI. These fully glycosylated full-length S protein model structures were assessed and further refined against the low-resoln. data in their resp. exptl. maps using ISOLDE. The authors then used CHARMM-GUI Membrane Builder to place the S proteins in a viral membrane and performed all-atom mol. dynamics simulations. All structures are available in CHARMM-GUI COVID-19 Archive so that researchers can use these models to carry out innovative and novel modeling and simulation research for the prevention and treatment of COVID-19.
- 37Casalino, L.; Gaieb, Z.; Goldsmith, J. A.; Hjorth, C. K.; Dommer, A. C.; Harbison, A. M.; Fogarty, C. A.; Barros, E. P.; Taylor, B. C.; McLellan, J. S. Beyond shielding: the roles of glycans in the SARS-CoV-2 spike protein. ACS Cent. Sci. 2020, 6, 1722– 1734, DOI: 10.1021/acscentsci.0c01056Google Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhvVOlsb3N&md5=52d499afcd7e3caa7d9e6017ffa86e45Beyond Shielding: The Roles of Glycans in the SARS-CoV-2 Spike ProteinCasalino, Lorenzo; Gaieb, Zied; Goldsmith, Jory A.; Hjorth, Christy K.; Dommer, Abigail C.; Harbison, Aoife M.; Fogarty, Carl A.; Barros, Emilia P.; Taylor, Bryn C.; McLellan, Jason S.; Fadda, Elisa; Amaro, Rommie E.ACS Central Science (2020), 6 (10), 1722-1734CODEN: ACSCII; ISSN:2374-7951. (American Chemical Society)The ongoing COVID-19 pandemic caused by severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) has resulted in more than 28,000,000 infections and 900,000 deaths worldwide to date. Antibody development efforts mainly revolve around the extensively glycosylated SARS-CoV-2 spike (S) protein, which mediates host cell entry by binding to the angiotensin-converting enzyme 2 (ACE2). Similar to many other viral fusion proteins, the SARS-CoV-2 spike utilizes a glycan shield to thwart the host immune response. Here, we built a full-length model of the glycosylated SARS-CoV-2 S protein, both in the open and closed states, augmenting the available structural and biol. data. Multiple microsecond-long, all-atom mol. dynamics simulations were used to provide an atomistic perspective on the roles of glycans and on the protein structure and dynamics. We reveal an essential structural role of N-glycans at sites N165 and N234 in modulating the conformational dynamics of the spike's receptor binding domain (RBD), which is responsible for ACE2 recognition. This finding is corroborated by biolayer interferometry expts., which show that deletion of these glycans through N165A and N234A mutations significantly reduces binding to ACE2 as a result of the RBD conformational shift toward the "down" state. Addnl., end-to-end accessibility analyses outline a complete overview of the vulnerabilities of the glycan shield of the SARS-CoV-2 S protein, which may be exploited in the therapeutic efforts targeting this mol. machine. Overall, this work presents hitherto unseen functional and structural insights into the SARS-CoV-2 S protein and its glycan coat, providing a strategy to control the conformational plasticity of the RBD that could be harnessed for vaccine development. The glycan shield is a sugary barrier that helps the viral SARS-CoV-2 spikes to evade the immune system. Beyond shielding, two of the spike's glycans are discovered to prime the virus for infection.
- 38Erbaş, A.; Netz, R. R. Confinement-dependent friction in peptide bundles. Biophys. J. 2013, 104, 1285– 1295, DOI: 10.1016/j.bpj.2013.02.008Google Scholar38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXktlKrs78%253D&md5=ccca5f67247a960798d12d28893b7797Confinement-Dependent Friction in Peptide BundlesErbas, Aykut; Netz, Roland R.Biophysical Journal (2013), 104 (6), 1285-1295CODEN: BIOJAU; ISSN:0006-3495. (Cell Press)Friction within globular proteins or between adhering macromols. crucially dets. the kinetics of protein folding, the formation, and the relaxation of self-assembled mol. systems. One fundamental question is how these friction effects depend on the local environment and in particular on the presence of water. In this model study, we use fully atomistic MD simulations with explicit water to obtain friction forces as a single polyglycine peptide chain is pulled out of a bundle of k adhering parallel polyglycine peptide chains. The whole system is periodically replicated along the peptide axes, so a stationary state at prescribed mean sliding velocity V is achieved. The aggregation no. is varied between k = 2 (two peptide chains adhering to each other with plenty of water present at the adhesion sites) and k = 7 (one peptide chain pulled out from a close-packed cylindrical array of six neighboring peptide chains with no water inside the bundle). The friction coeff. per hydrogen bond, extrapolated to the viscous limit of vanishing pulling velocity V → 0, exhibits an increase by five orders of magnitude when going from k = 2 to k = 7. This dramatic confinement-induced friction enhancement we argue to be due to a combination of water depletion and increased hydrogen-bond cooperativity.
- 39Erbaş, A.; Horinek, D.; Netz, R. R. Viscous friction of hydrogen-bonded matter. J. Am. Chem. Soc. 2012, 134, 623– 630, DOI: 10.1021/ja209454aGoogle Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhsV2jtb3K&md5=7fa0d24b3bd97c4821f9ba0f4c05aa3aViscous Friction of Hydrogen-Bonded MatterErbas, Aykut; Horinek, Dominik; Netz, Roland R.Journal of the American Chemical Society (2012), 134 (1), 623-630CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Amontons' law successfully describes friction between macroscopic solid bodies for a wide range of velocities and normal forces. For the diffusion and forced sliding of adhering or entangled macromols., proteins, and biol. complexes, temp. effects are invariably important, and a similarly successful friction law at biol. length and velocity scales is missing. Hydrogen bonds (HBs) are key to the specific binding of biomatter. Here we show that friction between hydrogen-bonded matter in the biol. relevant low-velocity viscous regime obeys a simple law: the friction force is proportional to the no. of HBs, the sliding velocity, and a friction coeff. γHB. This law is deduced from atomistic mol. dynamics simulations for short peptide chains that are laterally pulled over planar hydroxylated substrates in the presence of water and holds for widely different peptides, surface polarities, and applied normal forces. The value of γHB is extrapolated from simulations at sliding velocities in the range from V = 10-2 to 100 m/s by mapping on a simple stochastic model and turns out to be of the order of γHB ≃ 10-8 kg/s. The friction of a single HB thus amts. to the Stokes friction of a sphere with an equiv. radius of roughly 1 μm moving in water. Cooperativity is pronounced; roughly three HBs act collectively.
- 40Patil, S. P.; Xiao, S.; Gkagkas, K.; Markert, B.; Gräter, F. Viscous friction between crystalline and amorphous phase of dragline silk. PLoS One 2014, 9, e104832 DOI: 10.1371/journal.pone.0104832Google ScholarThere is no corresponding record for this reference.
- 41Kumar, S.; Rosenberg, J. M.; Bouzida, D.; Swendsen, R. H.; Kollman, P. A. The weighted histogram analysis method for free-energy calculations on biomolecules. I. The method. J. Comput. Chem. 1992, 13, 1011– 1021, DOI: 10.1002/jcc.540130812Google Scholar41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK38XmtVynsrs%253D&md5=5b2ad7410198f03025708a37c0fbe89dThe weighted histogram analysis method for free-energy calculations on biomolecules. I. The methodKumar, Shankar; Bouzida, Djamal; Swendsen, Robert H.; Kollman, Peter A.; Rosenberg, John M.Journal of Computational Chemistry (1992), 13 (8), 1011-21CODEN: JCCHDD; ISSN:0192-8651.The Weighted Histogram Anal. Method (WHAM), an extension of Ferrenberg and Swendsen's Multiple Histogram Technique, has been applied for the first time on complex biomol. Hamiltonians. The method is presented here as an extension of the Umbrella Sampling method for free-energy and Potential of Mean Force calcns. This algorithm possesses the following advantages over methods that are currently employed: (1) it provides a built-in est. of sampling errors thereby yielding objective ests. of the optimal location and length of addnl. simulations needed to achieve a desired level of precision; (2) it yields the "best" value of free energies by taking into account all the simulations so as to minimize the statistical errors; (3) in addn. to optimizing the links between simulations, it also allows multiple overlaps of probability distributions for obtaining better ests. of the free-energy differences. By recasting the Ferrenberg-Swendsen Multiple Histogram equations in a form suitable for mol. mechanics type Hamiltonians, we have demonstrated the feasibility and robustness of this method by applying it to a test problem of the generation of the Potential of Mean Force profile of the pseudorotation phase angle of the sugar ring in deoxyadenosine.
- 42Torrie, G. M.; Valleau, J. P. Nonphysical sampling distributions in Monte Carlo free-energy estimation: Umbrella sampling. J. Comput. Phys. 1977, 23, 187– 199, DOI: 10.1016/0021-9991(77)90121-8Google ScholarThere is no corresponding record for this reference.
- 43Phillips, R.; Kondev, J.; Theriot, J.; Garcia, H. Physical Biology of the Cell, 2nd ed.; Garland Science: New York, 2012; p 1088.Google ScholarThere is no corresponding record for this reference.
- 44Hanke, F.; Serr, A.; Kreuzer, H. J.; Netz, R. R. Stretching single polypeptides: The effect of rotational constraints in the backbone. Europhys. Lett. 2010, 92, 53001 DOI: 10.1209/0295-5075/92/53001Google ScholarThere is no corresponding record for this reference.
- 45Kitov, P. I.; Bundle, D. R. On the nature of the multivalency effect: a thermodynamic model. J. Am. Chem. Soc. 2003, 125, 16271– 16284, DOI: 10.1021/ja038223nGoogle Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXpsFOkuro%253D&md5=9ba6a3a7af88f5be5c790b32eb73aee1On the Nature of the Multivalency Effect: A Thermodynamic ModelKitov, Pavel I.; Bundle, David R.Journal of the American Chemical Society (2003), 125 (52), 16271-16284CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)A quant. model is proposed for the anal. of the thermodn. parameters of multivalent interactions in dil. solns. or with immobilized multimeric receptor. The model takes into account all bound species and describes multivalent binding via two microscopic binding energies corresponding to inter- and intramol. interactions (ΔG°inter and ΔG°intra), the relative contributions of which depend on the distribution of complexes with different nos. of occupied binding sites. The third component of the overall free energy, which we call the "avidity entropy" term, is a function of the degeneracy of bound states, Ωi, which is calcd. on the basis of the topol. of interaction and the distribution of all bound species. This term grows rapidly with the no. of receptor sites and ligand multivalency, it always favors binding, and explains why multivalency can overcome the loss of conformational entropy when ligands displayed at the ends of long tethers are bound. The microscopic parameters ΔG°inter and ΔG°intra may be detd. from the obsd. binding energies for a set of oligovalent ligands by nonlinear fitting with the theor. model. Here binding data obtained from two series of oligovalent carbohydrate inhibitors for Shiga-like toxins were used to verify the theory. The decavalent and octavalent inhibitors exhibit subnanomolar activity and are the most active sol. inhibitors yet seen that block Shiga-like toxin binding to its native receptor. The theory developed here in conjunction with our protocol for the optimization of tether length provides a predictive approach to design and maximize the avidity of multivalent ligands.
- 46Zumbro, E.; Witten, J.; Alexander-Katz, A. Computational insights into avidity of polymeric multivalent binders. Biophys. J. 2019, 117, 892– 902, DOI: 10.1016/j.bpj.2019.07.026Google Scholar46https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhsFemtb3L&md5=af2eb7445718e750eebf33397009c4aeComputational Insights into Avidity of Polymeric Multivalent BindersZumbro, Emiko; Witten, Jacob; Alexander-Katz, AlfredoBiophysical Journal (2019), 117 (5), 892-902CODEN: BIOJAU; ISSN:0006-3495. (Cell Press)Multivalent binding interactions are commonly found throughout biol. to enhance weak monovalent binding such as between glycoligands and protein receptors. Designing multivalent polymers to bind to viruses and toxic proteins is a promising avenue for inhibiting their attachment and subsequent infection of cells. Several studies have focused on oligomeric multivalent inhibitors and how changing parameters such as ligand shape, size, linker length, and flexibility affect binding. However, exptl. studies of how larger structural parameters of multivalent polymers, such as d.p., affect binding avidity to targets have mixed results, with some finding an improvement with longer polymers and some finding no effect. Here, we use Brownian dynamics simulations to provide a theor. understanding of how the d.p. affects the binding avidity of multivalent polymers. We show that longer polymers increase binding avidity to multivalent targets but reach a limit in binding avidity at high ds.p. We also show that when interacting with multiple targets simultaneously, longer polymers are able to use intertarget interactions to promote clustering and improve binding efficiency. We expect our results to narrow the design space for optimizing the structure and effectiveness of multivalent inhibitors as well as be useful to understand biol. design strategies for multivalent binding.
- 47Qiao, B.; Jiménez-Ángeles, F.; Nguyen, T. D.; Olvera de la Cruz, M. Water follows polar and nonpolar protein surface domains. Proc. Natl. Acad. Sci. U.S.A. 2019, 116, 19274– 19281, DOI: 10.1073/pnas.1910225116Google Scholar47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhvVGgu7jP&md5=cec3762bbfde8ac397491e088c567a9eWater follows polar and nonpolar protein surface domainsQiao, Baofu; Jimenez-Angeles, Felipe; Nguyen, Trung Dac; de la Cruz, Monica OlveraProceedings of the National Academy of Sciences of the United States of America (2019), 116 (39), 19274-19281CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)The conformation of water around proteins is of paramount importance, as it dets. protein interactions. Although the av. water properties around the surface of proteins have been provided exptl. and computationally, protein surfaces are highly heterogeneous. Therefore, it is crucial to det. the correlations of water to the local distributions of polar and nonpolar protein surface domains to understand functions such as aggregation, mutations, and delivery. By using atomistic simulations, we investigate the orientation and dynamics of water mols. next to 4 types of protein surface domains: neg. charged, pos. charged, and charge-neutral polar and nonpolar amino acids. The neg. charged amino acids orient around 98% of the neighboring water dipoles toward the protein surface, and such correlation persists up to around 16 Å from the protein surface. The pos. charged amino acids orient around 94% of the nearest water dipoles against the protein surface, and the correlation persists up to around 12 Å. The charge-neutral polar and nonpolar amino acids are also orienting the water neighbors in a quant. weaker manner. A similar trend was obsd. in the residence time of the nearest water neighbors. These findings hold true for 3 tech. important enzymes (PETase, cytochrome P 450, and organophosphorus hydrolase). Our results demonstrate that the water-amino acid degree of correlation follows the same trend as the amino acid contribution in proteins soly., namely, the neg. charged amino acids are the most beneficial for protein soly., then the pos. charged amino acids, and finally the charge-neutral amino acids.
- 48Manning, G. S. Limiting laws and counterion condensation in polyelectrolyte solutions I. Colligative properties. J. Chem. Phys. 1969, 51, 924– 933, DOI: 10.1063/1.1672157Google Scholar48https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaF1MXkslCrsLk%253D&md5=9e5de22424890b7e855a638456724ddfLimiting laws and counterion condensation in polyelectroyte solutions. I. Colligative propertiesManning, Gerald S.Journal of Chemical Physics (1969), 51 (3), 924-33CODEN: JCPSA6; ISSN:0021-9606.Formulas are derived for the osmotic coeff., the Donnan salt-exclusion factor, and the mobile-ion activity coeffs. in a polyelectrolyte soln. with or without added sample salt. The formulas, which contain no adjustable parameters, are based on the (theoretical) observation by several workers that counterions will "condense" on the polyion until the charge d. on the polyion is reduced below a certain crit. value. The uncondensed mobile ions are treated in the Debye-Hueckel approxn. In a restricted sense, the formulas are "limiting laws," and this aspect is discussed at length. Detailed comparison with exptl. data in the literature is given; agreement of the theory with expt. is usually quant.
- 49Fenley, M. O.; Manning, G. S.; Olson, W. K. Approach to the limit of counterion condensation. Biopolymers 1990, 30, 1191– 1203, DOI: 10.1002/bip.360301305Google Scholar49https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3MXhvVSmsLg%253D&md5=0370802614aafca9380719feb02c39a1Approach to the limit of counterion condensationFenley, Marcia O.; Manning, Gerald S.; Olson, Wilma K.Biopolymers (1990), 30 (13-14), 1191-203CODEN: BIPMAA; ISSN:0006-3525.According to counterion condensation theory, one of the contributions to the polyelectrolyte free energy is a pairwise sum of Debye-Hueckel potentials between polymer charges that are reduced by condensed counterions. When the polyion model is taken as an infinitely long and uniformly spaced line of charges, a simple closed expression for the summation, combined with entropy-derived mixing contributions, leads to the central result of the theory, a condensed fraction of counterions dependent only on the linear charge d. of the polyion and the valence of the counterion, stable against increases of salt up to concns. in excess of 0.1M. Here the sum is numerically evaluated for B-DNA models other than the infinite line of B-DNA charges. For a finite-length line there are end effects at low salt. The condensation limit is reached as a flat plateau by increasing the salt concn. At a fixed salt concn. the condensation limit is reached by increasing the length of the line. At moderate salt even very short B-DNA line-model oligomers have condensed fractions not far from the infinite polymer limit. For a long double-helical array with charge coordinates at the phosphates of B-DNA, the limiting condensed fraction appears to be approached at low salt. In contrast to the results for the line of charges, however, the computed condensed fraction varies strongly with salt in the range of exptl. typical concns. Salt invariance is restored, in agreement with both the line model and exptl. data, when dielec. satn. is considered by means of a distance-dependent dielec. function. For sufficiently long B-DNA line and helical models, at typical salt concns., the counterion binding fraction approaches the polymer limit as a linear function of 1/P, where P is the no. of phosphate groups of B-DNA.
- 50Walkowiak, J. J.; Ballauff, M. Interaction of polyelectrolytes with proteins: quantifying the role of water. Adv. Sci. 2021, 8, 2100661 DOI: 10.1002/advs.202100661Google Scholar50https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXisV2qu7bM&md5=78e86f5b8efd62947c1f047b2890347fInteraction of Polyelectrolytes with Proteins: Quantifying the Role of WaterWalkowiak, Jacek J.; Ballauff, MatthiasAdvanced Science (Weinheim, Germany) (2021), 8 (12), 2100661CODEN: ASDCCF; ISSN:2198-3844. (Wiley-VCH Verlag GmbH & Co. KGaA)A theor. model is presented for the free energy ΔGb of complex formation between a highly charged polyelectrolyte and a protein. The model introduced here comprises both the effect of released counterions and the uptake or release of water mols. during complex formation. The resulting expression for ΔGb is hence capable of describing the dependence of ΔGb on temp. as well as on the concn. of salt in the system: An increase of the salt concn. in the soln. increases the activity of the ions and counterion release becomes less effective for binding. On the other hand, an increased salt concn. leads to the decrease of the activity of water in bulk. Hence, release of water mols. during complex formation will be more advantageous and lead to an increase of the magnitude of ΔGb and the binding const. It is furthermore demonstrated that the release or uptake of water mols. is the origin of the marked enthalpy-entropy cancellation obsd. during complex formation of polyelectrolytes with proteins. The comparison with exptl. data on complex formation between a synthetic (sulfated dendritic polyglycerol) and natural polyelectrolytes (DNA; heparin) with proteins shows full agreement with theory.
- 51Irudayam, S. J.; Henchman, R. H. Entropic cost of protein- ligand binding and its dependence on the entropy in solution. J. Phys. Chem. B 2009, 113, 5871– 5884, DOI: 10.1021/jp809968pGoogle Scholar51https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXktVOltLY%253D&md5=997977e2ace284e3af8e6f46ae347dd0Entropic Cost of Protein-Ligand Binding and Its Dependence on the Entropy in SolutionIrudayam, Sheeba Jem; Henchman, Richard H.Journal of Physical Chemistry B (2009), 113 (17), 5871-5884CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)Two theor. formulations are proposed and compared for the loss of translational and rotational entropy upon protein-ligand binding in water. The two theories share the same approach to evaluate the translational and rotational entropy of the ligand when bound. The potential of the bound ligand is modeled by six harmonic oscillators that are parametrized from the force and torque magnitudes measured in a mol. dynamics simulation, yielding vibrational and librational entropies. In the aq. phase, the theories differ because there is no unique way to assign the total entropy to mols. in soln. In one approach, the ligand is allowed unrestricted access to the full soln. vol. at the std. concn. and is assigned the same translational and rotational entropy as if it were an ideal gas. We term this a "mol.-frame" (MF) theory because it considers configurational space in the ref. frame of the mol. of interest. The entropy of the solvent is penalized because it is excluded from the mol.'s vol. In the second theory, all mols. including the solvent are confined by their neighbors in mean-field configurational vols. This we term a "system-frame" (SF) theory because the configurational space available to all mols. is considered in the ref. frame of the whole system. Mols. have vibrational and librational entropy in the same way as they do when bound. In addn., the discrete size of the solvent mols. quantizes the configurational space into an effective no. of min. according to the solute mol.'s std. concn. and the mean vol. of a solvent mol. This leads to the cratic entropy expressed in terms of the solute mol.'s mole fraction. The equivalent no. of min. in rotational space depends on both the solute mol.'s vol. and the solvent mol.'s vol. This leads to an equation for the orientational entropy based on the proposed concept of "angle fraction". The MF and SF theories are applied to calc. the translational and rotational entropy losses involved in the formation of six different protein-ligand complexes, in two of which the ligand is water. The MF entropy losses range from -80 to -142 J K-1 mol-1 for ligands at the 1 M std.-state concn. and from -52 to -63 J K-1 mol-1 for water at the 55.6 M std.-state concn. They depend logarithmically on both the no. and strength of interactions between the ligand and protein through the forces and torques. This is obsd. to lead to moderate dependencies on the ligands' moments of inertia and masses. The SF entropy losses are smaller and range from -50 to -75 J K-1 mol-1 for ligands at the 1 M std.-state concn. and from 0 to -12 J K-1 mol-1 for water. They depend logarithmically on the ligand solvent's mol. vol. and weakly on the relative strengths of the ligand's interactions with the protein and water. The cratic entropy loss in water at the std. concn. is const. and is also demonstrated to be implicit in MF theories. Entropy losses from the two approaches are also compared with those from other computational approaches and with expt. The use of the force and torque magnitudes leads to smaller bound vols. than are obtained from ligand-displacement approaches. The general agreement of the SF entropy losses with those from expt. suggests that the SF theory is more consistent with the assumptions made in exptl. measurements than the MF solvation theories, which would require a compensating entropy gain in the solvent in order to agree.
- 52Xu, X.; Ran, Q.; Dey, P.; Nikam, R.; Haag, R.; Ballauff, M.; Dzubiella, J. Counterion-release entropy governs the inhibition of serum proteins by polyelectrolyte drugs. Biomacromolecules 2018, 19, 409– 416, DOI: 10.1021/acs.biomac.7b01499Google Scholar52https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXitVWgurvN&md5=82e6f191e2e59bdd495800992841eca1Counterion-Release Entropy Governs the Inhibition of Serum Proteins by Polyelectrolyte DrugsXu, Xiao; Ran, Qidi; Dey, Pradip; Nikam, Rohit; Haag, Rainer; Ballauff, Matthias; Dzubiella, JoachimBiomacromolecules (2018), 19 (2), 409-416CODEN: BOMAF6; ISSN:1525-7797. (American Chemical Society)Dendritic polyelectrolytes constitute high potential drugs and carrier systems for biomedical purposes. Still, their biomol. interaction modes, in particular those detg. the binding affinity to proteins, have not been rationalized. We study the interaction of the drug candidate dendritic polyglycerol sulfate (dPGS) with serum proteins using isothermal titrn. calorimetry (ITC) interpreted and complemented with mol. computer simulations. Lysozyme is first studied as a well-defined model protein to verify theor. concepts, which are then applied to the important cell adhesion protein family of selectins. We demonstrate that the driving force of the strong complexation, leading to a distinct protein corona, originates mainly from the release of only a few condensed counterions from the dPGS upon binding. The binding const. shows a surprisingly weak dependence on dPGS size (and bare charge) which can be understood by colloidal charge-renormalization effects and by the fact that the magnitude of the dominating counterion-release mechanism almost exclusively depends on the interfacial charge structure of the protein-specific binding patch. Our findings explain the high selectivity of P- and L-selectins over E-selectin for dPGS to act as a highly anti-inflammatory drug. The entire anal. demonstrates that the interaction of proteins with charged polymeric drugs can be predicted by simulations with unprecedented accuracy. Thus, our results open new perspectives for the rational design of charged polymeric drugs and carrier systems.
- 53Caro, J. A.; Harpole, K. W.; Kasinath, V.; Lim, J.; Granja, J.; Valentine, K. G.; Sharp, K. A.; Wand, A. J. Entropy in molecular recognition by proteins. Proc. Natl. Acad. Sci. U.S.A. 2017, 114, 6563– 6568, DOI: 10.1073/pnas.1621154114Google Scholar53https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXpt1Onsbo%253D&md5=7fc266b399633ec8d63e12bce098e228Entropy in molecular recognition by proteinsCaro, Jose A.; Harpole, Kyle W.; Kasinath, Vignesh; Lim, Jackwee; Granja, Jeffrey; Valentine, Kathleen G.; Sharp, Kim A.; Wand, A. JoshuaProceedings of the National Academy of Sciences of the United States of America (2017), 114 (25), 6563-6568CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)Mol. recognition by proteins is fundamental to mol. biol. Dissection of the thermodn. energy terms governing protein-ligand interactions has proven difficult, with detn. of entropic contributions being particularly elusive. NMR relaxation measurements have suggested that changes in protein conformational entropy can be quant. obtained through a dynamical proxy, but the generality of this relationship has not been shown. Twenty-eight protein-ligand complexes are used to show a quant. relationship between measures of fast side-chain motion and the underlying conformational entropy. We find that the contribution of conformational entropy can range from favorable to unfavorable, which demonstrates the potential of this thermodn. variable to modulate protein-ligand interactions. For about one-quarter of these complexes, the absence of conformational entropy would render the resulting affinity biol. meaningless. The dynamical proxy for conformational entropy or "entropy meter" also allows for refinement of the contributions of solvent entropy and the loss in rotational-translational entropy accompanying formation of high-affinity complexes. Furthermore, structure-based application of the approach can also provide insight into long-lived specific water-protein interactions that escape the generic treatments of solvent entropy based simply on changes in accessible surface area. These results provide a comprehensive and unified view of the general role of entropy in high-affinity mol. recognition by proteins.
- 54Sahoo, A. K.; Schreiber, F.; Netz, R. R.; Maiti, P. K. Role of entropy in determining the phase behavior of protein solutions induced by multivalent ions. Soft Matter 2022, 18, 592– 601, DOI: 10.1039/D1SM00730KGoogle ScholarThere is no corresponding record for this reference.
- 55Rapaport, D. Hydrogen bonds in water: Network organization and lifetimes. Mol. Phys. 1983, 50, 1151– 1162, DOI: 10.1080/00268978300102931Google Scholar55https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2cXhvVSntrw%253D&md5=9a83628aacab20bcca277cd3ae17c203Hydrogen bonds in water. Network organization and lifetimesRapaport, D. C.Molecular Physics (1983), 50 (5), 1151-62CODEN: MOPHAM; ISSN:0026-8976.Equil. and dynamical properties of H bonds in liq. H2O are analyzed using the results of mol. dynamics simulations of the MCY-CI model. Properties of the H bond clusters as functions of temp. are described. The connectivity of the clusters is analyzed in terms of the bridgeless polygons which are formed by the bonds. The problem of obtaining a meaningful definition of bond lifetime is discussed, and the results of lifetime measurements based on alternative definitions are shown.
- 56Malicka, W.; Haag, R.; Ballauff, M. Interaction of heparin with proteins: hydration effects. J. Phys. Chem. B 2022, 126, 6250– 6260, DOI: 10.1021/acs.jpcb.2c04928Google ScholarThere is no corresponding record for this reference.
- 57Geisler, M.; Xiao, S.; Puchner, E. M.; Gräter, F.; Hugel, T. Controlling the structure of proteins at surfaces. J. Am. Chem. Soc. 2010, 132, 17277– 17281, DOI: 10.1021/ja107212zGoogle Scholar57https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhsVeksL%252FI&md5=c7af30feab283481a7e018b07bc0892eControlling the Structure of Proteins at SurfacesGeisler, Michael; Xiao, Sen-Bo; Puchner, Elias M.; Graeter, Frauke; Hugel, ThorstenJournal of the American Chemical Society (2010), 132 (48), 17277-17281CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)With the help of single mol. force spectroscopy and mol. dynamics simulations, we det. the surface-induced structure of a single engineered spider silk protein. An amyloid like structure is induced in the vicinity of a surface with high surface energy and can be prohibited in the presence of a hydrophobic surface. The derived mol. energy landscapes highlight the role of single silk protein structure for the macroscopic toughness of spider silk.
- 58Liese, S.; Netz, R. R. Influence of length and flexibility of spacers on the binding affinity of divalent ligands. Beilstein J. Org. Chem. 2015, 11, 804– 816, DOI: 10.3762/bjoc.11.90Google Scholar58https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXpvFGntro%253D&md5=43e11f85192cf5716dafcc802eec77d3Influence of length and flexibility of spacers on the binding affinity of divalent ligandsLiese, Susanne; Netz, Roland R.Beilstein Journal of Organic Chemistry (2015), 11 (), 804-816CODEN: BJOCBH; ISSN:1860-5397. (Beilstein-Institut zur Foerderung der Chemischen Wissenschaften)We present a quant. model for the binding of divalent ligand-receptor systems. We study the influence of length and flexibility of the spacers on the overall binding affinity and derive general rules for the optimal ligand design. To this end, we first compare different polymeric models and det. the probability to simultaneously bind to two neighboring receptor binding pockets. In a second step the binding affinity of divalent ligands in terms of the IC50 value is derived. We find that a divalent ligand has the potential to bind more efficiently than its monovalent counterpart only, if the monovalent dissocn. const. is lower than a crit. value. This crit. monovalent dissocn. const. depends on the ligand-spacer length and flexibility as well as on the size of the receptor. Regarding the optimal ligand-spacer length and flexibility, we find that the av. spacer length should be equal or slightly smaller than the distance between the receptor binding pockets and that the end-to-end spacer length fluctuations should be in the same range as the size of a receptor binding pocket.
- 59Jo, S.; Kim, T.; Iyer, V. G.; Im, W. CHARMM-GUI: a web-based graphical user interface for CHARMM. J. Comput. Chem. 2008, 29, 1859– 1865, DOI: 10.1002/jcc.20945Google Scholar59https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXosVKksbc%253D&md5=112a3dd61d792b040f9f716b32220d7eCHARMM-GUI: a web-based graphical user interface for CHARMMJo, Sunhwan; Kim, Taehoon; Iyer, Vidyashankara G.; Im, WonpilJournal of Computational Chemistry (2008), 29 (11), 1859-1865CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)CHARMM is an academic research program used widely for macromol. mechanics and dynamics with versatile anal. and manipulation tools of at. coordinates and dynamics trajectories. CHARMM-GUI, http://www.charmm-gui.org, has been developed to provide a web-based graphical user interface to generate various input files and mol. systems to facilitate and standardize the usage of common and advanced simulation techniques in CHARMM. The web environment provides an ideal platform to build and validate a mol. model system in an interactive fashion such that, if a problem is found through visual inspection, one can go back to the previous setup and regenerate the whole system again. In this article, we describe the currently available functional modules of CHARMM-GUI Input Generator that form a basis for the advanced simulation techniques. Future directions of the CHARMM-GUI development project are also discussed briefly together with other features in the CHARMM-GUI website, such as Archive and Movie Gallery.
- 60Park, S.-J.; Lee, J.; Qi, Y.; Kern, N. R.; Lee, H. S.; Jo, S.; Joung, I.; Joo, K.; Lee, J.; Im, W. CHARMM-GUI Glycan Modeler for modeling and simulation of carbohydrates and glycoconjugates. Glycobiology 2019, 29, 320– 331, DOI: 10.1093/glycob/cwz003Google Scholar60https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXitFajsLzJ&md5=0bd9d6b408efabd1d1a1440c9a639100CHARMM-GUI Glycan Modeler for modeling and simulation of carbohydrates and glycoconjugatesPark, Sang-Jun; Lee, Jumin; Qi, Yifei; Kern, Nathan R.; Lee, Hui Sun; Jo, Sunhwan; Joung, Insuk; Joo, Keehyung; Lee, Jooyoung; Im, WonpilGlycobiology (2019), 29 (4), 320-331CODEN: GLYCE3; ISSN:1460-2423. (Oxford University Press)Characterizing glycans and glycoconjugates in context of three-dimensional structures is important in understanding their biol. roles and developing efficient therapeutic agents. Computational modeling and mol. simulation have become essential tool complementary to exptl. methods. Here, we present computational tool, Glycan Modeler for in silico N-/O-glycosylation of target protein and generation of carbohydrate-only systems. In our previous study, we developed Glycan Reader, web-based tool for detecting carbohydrate mols. from PDB structure and generation of simulation system and input files. As integrated into Glycan Reader in CHARMM-GUI, Glycan Modeler enables to generate structures of glycans and glycoconjugates for given glycan sequences and glycosylation sites using PDB glycan template structures from Glycan Fragment Database (http://glycanstructure.org/ fragment-db). Our benchmark tests demonstrate universal applicability of Glycan Reader & Modeler to various glycan sequences and target proteins. We also investigated structural properties of modeled glycan structures by running 2-μs mol. dynamics simulations of HIV envelope protein. Simulations show that modeled glycan structures built by Glycan Reader & Modeler have similar structural features compared to ones solved by X-ray crystallog. We also describe representative examples of glycoconjugate modeling with video demos to illustrate practical applications of Glycan Reader & Modeler.
- 61Jo, S.; Song, K. C.; Desaire, H.; MacKerell, A. D., Jr; Im, W. Glycan Reader: automated sugar identification and simulation preparation for carbohydrates and glycoproteins. J. Comput. Chem. 2011, 32, 3135– 3141, DOI: 10.1002/jcc.21886Google Scholar61https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXpvVamtrk%253D&md5=0cd04c0fb892d58ae69501f23ab079dfGlycan Reader: Automated sugar identification and simulation preparation for carbohydrates and glycoproteinsJo, Sunhwan; Song, Kevin C.; Desaire, Heather; MacKerell, Alexander D.; Im, WonpilJournal of Computational Chemistry (2011), 32 (14), 3135-3141CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)Understanding how glycosylation affects protein structure, dynamics, and function is an emerging and challenging problem in biol. As a first step toward glycan modeling in the context of structural glycobiol., the authors have developed Glycan Reader and integrated it into the CHARMM-GUI,. Glycan Reader greatly simplifies the reading of PDB structure files contg. glycans through (i) detection of carbohydrate mols., (ii) automatic annotation of carbohydrates based on their three-dimensional structures, (iii) recognition of glycosidic linkages between carbohydrates as well as N-/O-glycosidic linkages to proteins, and (iv) generation of inputs for the biomol. simulation program CHARMM with the proper glycosidic linkage setup. In addn., Glycan Reader is linked to other functional modules in CHARMM-GUI, allowing users to easily generate carbohydrate or glycoprotein mol. simulation systems in soln. or membrane environments and visualize the electrostatic potential on glycoprotein surfaces. These tools are useful for studying the impact of glycosylation on protein structure and dynamics. © 2011 Wiley Periodicals, Inc. J Comput Chem, 2011.
- 62Park, S.-J.; Lee, J.; Patel, D. S.; Ma, H.; Lee, H. S.; Jo, S.; Im, W. Glycan Reader is improved to recognize most sugar types and chemical modifications in the Protein Data Bank. Bioinformatics 2017, 33, 3051– 3057, DOI: 10.1093/bioinformatics/btx358Google Scholar62https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhvFGju7bO&md5=16900496c68d5ffeedb08ed0e67e0b94Glycan Reader is improved to recognize most sugar types and chemical modifications in the Protein Data BankPark, Sang-Jun; Lee, Jumin; Patel, Dhilon S.; Ma, Hongjing; Lee, Hui Sun; Jo, Sunhwan; Im, WonpilBioinformatics (2017), 33 (19), 3051-3057CODEN: BOINFP; ISSN:1367-4811. (Oxford University Press)Motivation: Glycans play a central role in many essential biol. processes. Glycan Reader was originally developed to simplify the reading of Protein Data Bank (PDB) files contg. glycans through the automatic detection and annotation of sugars and glycosidic linkages between sugar units and to proteins, all based on at. coordinates and connectivity information. Carbohydrates can have various chem. modifications at different positions, making their chem. space much diverse. Unfortunately, current PDB files do not provide exact annotations for most carbohydrate derivs. and more than 50% of PDB glycan chains have at least one carbohydrate deriv. that could not be correctly recognized by the original Glycan Reader. Results: Glycan Reader has been improved and now identifies most sugar types and chem. modifications (including various glycolipids) in the PDB, and both PDB and PDBx/mmCIF formats are supported. CHARMM-GUI Glycan Reader is updated to generate the simulation system and input of various glycoconjugates with most sugar types and chem. modifications. It also offers a new functionality to edit the glycan structures through addn./deletion/modification of glycosylation types, sugar types, chem. modifications, glycosidic linkages, and anomeric states. The simulation system and input files can be used for CHARMM, NAMD, GROMACS, AMBER, GENESIS, LAMMPS, Desmond, OpenMM, and CHARMM/OpenMM. Glycan Fragment Database in GlycanStructure.Org is also updated to provide an intuitive glycan sequence search tool for complex glycan structures with various chem. modifications in the PDB.
- 63Hanwell, M. D.; Curtis, D. E.; Lonie, D. C.; Vandermeersch, T.; Zurek, E.; Hutchison, G. R. Avogadro: an advanced semantic chemical editor, visualization, and analysis platform. J. Cheminf. 2012, 4, 1– 17, DOI: 10.1186/1758-2946-4-17Google ScholarThere is no corresponding record for this reference.
- 64Huang, J.; Rauscher, S.; Nawrocki, G.; Ran, T.; Feig, M.; De Groot, B. L.; Grubmüller, H.; MacKerell, A. D. CHARMM36m: an improved force field for folded and intrinsically disordered proteins. Nat. Methods 2017, 14, 71– 73, DOI: 10.1038/nmeth.4067Google Scholar64https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhvVSiu77I&md5=0aa151fbef2ee0b5e2cfb593c54330c2CHARMM36m: an improved force field for folded and intrinsically disordered proteinsHuang, Jing; Rauscher, Sarah; Nawrocki, Grzegorz; Ran, Ting; Feig, Michael; de Groot, Bert L.; Grubmuller, Helmut; MacKerell, Alexander D. JrNature Methods (2017), 14 (1), 71-73CODEN: NMAEA3; ISSN:1548-7091. (Nature Publishing Group)The all-atom additive CHARMM36 protein force field is widely used in mol. modeling and simulations. We present its refinement, CHARMM36m (http://mackerell.umaryland.edu/charmm_ff.shtml), with improved accuracy in generating polypeptide backbone conformational ensembles for intrinsically disordered peptides and proteins.
- 65Guvench, O.; Mallajosyula, S. S.; Raman, E. P.; Hatcher, E.; Vanommeslaeghe, K.; Foster, T. J.; Jamison, F. W.; MacKerell, A. D., Jr CHARMM additive all-atom force field for carbohydrate derivatives and its utility in polysaccharide and carbohydrate-protein modeling. J. Chem. Theory Comput. 2011, 7, 3162– 3180, DOI: 10.1021/ct200328pGoogle Scholar65https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtFSru77P&md5=fc3c05ef4c95895ee3e7b08472ef3554CHARMM Additive All-Atom Force Field for Carbohydrate Derivatives and Its Utility in Polysaccharide and Carbohydrate-Protein ModelingGuvench, Olgun; Mallajosyula, Sairam S.; Raman, E. Prabhu; Hatcher, Elizabeth; Vanommeslaeghe, Kenno; Foster, Theresa J.; Jamison, Francis W.; MacKerell, Alexander D.Journal of Chemical Theory and Computation (2011), 7 (10), 3162-3180CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Monosaccharide derivs. such as xylose, fucose, N-acetylglucosamine (GlcNAc), N-acetylgalactosamine (GlaNAc), glucuronic acid, iduronic acid, and N-acetylneuraminic acid (Neu5Ac) are important components of eukaryotic glycans. The present work details the development of force-field parameters for these monosaccharides and their covalent connections to proteins via O linkages to serine or threonine side chains and via N linkages to asparagine side chains. The force field development protocol was designed to explicitly yield parameters that are compatible with the existing CHARMM additive force field for proteins, nucleic acids, lipids, carbohydrates, and small mols. Therefore, when combined with previously developed parameters for pyranose and furanose monosaccharides, for glycosidic linkages between monosaccharides, and for proteins, the present set of parameters enables the mol. simulation of a wide variety of biol. important mols. such as complex carbohydrates and glycoproteins. Parametrization included fitting to quantum mech. (QM) geometries and conformational energies of model compds., as well as to QM pair interaction energies and distances of model compds. with water. Parameters were validated in the context of crystals of relevant monosaccharides, as well NMR and/or x-ray crystallog. data on larger systems including oligomeric hyaluronan, sialyl Lewis X, O- and N-linked glycopeptides, and a lectin:sucrose complex. As the validated parameters are an extension of the CHARMM all-atom additive biomol. force field, they further broaden the types of heterogeneous systems accessible with a consistently developed force-field model.
- 66Mallajosyula, S. S.; Guvench, O.; Hatcher, E.; MacKerell, A. D., Jr CHARMM additive all-atom force field for phosphate and sulfate linked to carbohydrates. J. Chem. Theory Comput. 2012, 8, 759– 776, DOI: 10.1021/ct200792vGoogle Scholar66https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhs1Oju7zP&md5=8350bcdedba6a5e80675f9cae77dfc7eCHARMM Additive All-Atom Force Field for Phosphate and Sulfate Linked to CarbohydratesMallajosyula, Sairam S.; Guvench, Olgun; Hatcher, Elizabeth; MacKerell, Alexander D.Journal of Chemical Theory and Computation (2012), 8 (2), 759-776CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Presented is an extension of the CHARMM additive all-atom carbohydrate force field to enable the modeling of phosphate and sulfate linked to carbohydrates. The parameters are developed in a hierarchical fashion using model compds. contg. the key atoms in the full carbohydrates. Target data for parameter optimization included full two-dimensional energy surfaces defined by the glycosidic dihedral angle pairs in the phosphate/sulfate model compd. analogs of hexopyranose monosaccharide phosphates and sulfates, as detd. by quantum mech. (QM) MP2/cc-pVTZ single point energies on MP2/6-31+G(d) optimized structures. To achieve balanced, transferable dihedral parameters for the dihedral angles, surfaces for all possible anomeric and conformational states were included during the parametrization process. To model physiol. relevant systems, both the mono- and dianionic charged states were studied for the phosphates. This resulted in over 7000 MP2/cc-pVTZ//MP2/6-31G+(d) model compd. conformational energies which, supplemented with QM geometries, were the main target data for the parametrization. Parameters were validated against crystals of relevant monosaccharide derivs. obtained from the Cambridge Structural Database (CSD) and larger systems, inositol-(tri/tetra/penta) phosphates noncovalently bound to the pleckstrin homol. (PH) domain and oligomeric chondroitin sulfate in soln. and in complex with cathepsin K protein.
- 67Vanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.; Mackerell, A. D. CHARMM general force field: A force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fields. J. Comput. Chem. 2010, 31, 671– 690, DOI: 10.1002/jcc.21367Google Scholar67https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhtlentbc%253D&md5=26212e0e4f73bded0c89d4b411cd3833CHARMM general force field: A force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fieldsVanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.; Mackerell, A. D., Jr.Journal of Computational Chemistry (2010), 31 (4), 671-690CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)The widely used CHARMM additive all-atom force field includes parameters for proteins, nucleic acids, lipids, and carbohydrates. In the present article, an extension of the CHARMM force field to drug-like mols. is presented. The resulting CHARMM General Force Field (CGenFF) covers a wide range of chem. groups present in biomols. and drug-like mols., including a large no. of heterocyclic scaffolds. The parametrization philosophy behind the force field focuses on quality at the expense of transferability, with the implementation concg. on an extensible force field. Statistics related to the quality of the parametrization with a focus on exptl. validation are presented. Addnl., the parametrization procedure, described fully in the present article in the context of the model systems, pyrrolidine, and 3-phenoxymethyl-pyrrolidine will allow users to readily extend the force field to chem. groups that are not explicitly covered in the force field as well as add functional groups to and link together mols. already available in the force field. CGenFF thus makes it possible to perform "all-CHARMM" simulations on drug-target interactions thereby extending the utility of CHARMM force fields to medicinally relevant systems. © 2009 Wiley Periodicals, Inc.J Comput Chem, 2010.
- 68Yu, W.; He, X.; Vanommeslaeghe, K.; MacKerell, A. D., Jr Extension of the CHARMM general force field to sulfonyl-containing compounds and its utility in biomolecular simulations. J. Comput. Chem. 2012, 33, 2451– 2468, DOI: 10.1002/jcc.23067Google Scholar68https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhtVKntL7J&md5=8de6f9a92b8ad706b010534a51fe54cdExtension of the CHARMM general force field to sulfonyl-containing compounds and its utility in biomolecular simulationsYu, Wenbo; He, Xibing; Vanommeslaeghe, Kenno; MacKerell, Alexander D.Journal of Computational Chemistry (2012), 33 (31), 2451-2468CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)Presented is an extension of the CHARMM General Force Field (CGenFF) to enable the modeling of sulfonyl-contg. compds. Model compds. contg. chem. moieties such as sulfone, sulfonamide, sulfonate, and sulfamate were used as the basis for the parameter optimization. Targeting high-level quantum mech. and exptl. crystal data, the new parameters were optimized in a hierarchical fashion designed to maintain compatibility with the remainder of the CHARMM additive force field. The optimized parameters satisfactorily reproduced equil. geometries, vibrational frequencies, interactions with water, gas phase dipole moments, and dihedral potential energy scans. Validation involved both cryst. and liq. phase calcns. showing the newly developed parameters to satisfactorily reproduce exptl. unit cell geometries, crystal intramol. geometries, and pure solvent densities. The force field was subsequently applied to study conformational preference of a sulfonamide-based peptide system. Good agreement with exptl. IR/NMR data further validated the newly developed CGenFF parameters as a tool to investigate the dynamic behavior of sulfonyl groups in a biol. environment. CGenFF now covers sulfonyl group contg. moieties allowing for modeling and simulation of sulfonyl-contg. compds. in the context of biomol. systems including compds. of medicinal interest.
- 69Vanommeslaeghe, K.; MacKerell, A. D., Jr Automation of the CHARMM General Force Field (CGenFF) I: bond perception and atom typing. J. Chem. Inf. Model. 2012, 52, 3144– 3154, DOI: 10.1021/ci300363cGoogle Scholar69https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhs1Gns7fL&md5=c6679293f4a2501f2bcadf2020ca1473Automation of the CHARMM General Force Field (CGenFF) I: Bond Perception and Atom TypingVanommeslaeghe, K.; MacKerell, A. D.Journal of Chemical Information and Modeling (2012), 52 (12), 3144-3154CODEN: JCISD8; ISSN:1549-9596. (American Chemical Society)Mol. mechanics force fields are widely used in computer-aided drug design for the study of drug-like mols. alone or interacting with biol. systems. In simulations involving biol. macromols., the biol. part is typically represented by a specialized biomol. force field, while the drug is represented by a matching general (org.) force field. In order to apply these general force fields to an arbitrary drug-like mol., functionality for assignment of atom types, parameters, and charges is required. In the present article, which is part I of a series of two, we present the algorithms for bond perception and atom typing for the CHARMM General Force Field (CGenFF). The CGenFF atom typer first assocs. attributes to the atoms and bonds in a mol., such as valence, bond order, and ring membership among others. Of note are a no. of features that are specifically required for CGenFF. This information is then used by the atom typing routine to assign CGenFF atom types based on a programmable decision tree. This allows for straight-forward implementation of CGenFF's complicated atom typing rules and for equally straight-forward updating of the atom typing scheme as the force field grows. The presented atom typer was validated by assigning correct atom types on 477 model compds. including in the training set as well as 126 test-set mols. that were constructed to specifically verify its different components. The program may be utilized via an online implementation at https://www.paramchem.org/.
- 70Vanommeslaeghe, K.; Raman, E. P.; MacKerell, A. D., Jr Automation of the CHARMM General Force Field (CGenFF) II: assignment of bonded parameters and partial atomic charges. J. Chem. Inf. Model. 2012, 52, 3155– 3168, DOI: 10.1021/ci3003649Google Scholar70https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhs1Gns7fF&md5=e676ad1f42cb1e98dd353d4d285e8d13Automation of the CHARMM General Force Field (CGenFF) II: Assignment of Bonded Parameters and Partial Atomic ChargesVanommeslaeghe, K.; Raman, E. Prabhu; MacKerell, A. D.Journal of Chemical Information and Modeling (2012), 52 (12), 3155-3168CODEN: JCISD8; ISSN:1549-9596. (American Chemical Society)Mol. mechanics force fields are widely used in computer-aided drug design for the study of drug candidates interacting with biol. systems. In these simulations, the biol. part is typically represented by a specialized biomol. force field, while the drug is represented by a matching general (org.) force field. In order to apply these general force fields to an arbitrary drug-like mol., functionality for assignment of atom types, parameters, and partial at. charges is required. In the present article, algorithms for the assignment of parameters and charges for the CHARMM General Force Field (CGenFF) are presented. These algorithms rely on the existing parameters and charges that were detd. as part of the parametrization of the force field. Bonded parameters are assigned based on the similarity between the atom types that define said parameters, while charges are detd. using an extended bond-charge increment scheme. Charge increments were optimized to reproduce the charges on model compds. that were part of the parametrization of the force field. Case studies are presented to clarify the functioning of the algorithms and the significance of their output data.
- 71Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R. W.; Klein, M. L. Comparison of simple potential functions for simulating liquid water. J. Chem. Phys. 1983, 79, 926– 935, DOI: 10.1063/1.445869Google Scholar71https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3sXksF2htL4%253D&md5=a1161334e381746be8c9b15a5e56f704Comparison of simple potential functions for simulating liquid waterJorgensen, William L.; Chandrasekhar, Jayaraman; Madura, Jeffry D.; Impey, Roger W.; Klein, Michael L.Journal of Chemical Physics (1983), 79 (2), 926-35CODEN: JCPSA6; ISSN:0021-9606.Classical Monte Carlo simulations were carried out for liq. H2O in the NPT ensemble at 25° and 1 atm using 6 of the simpler intermol. potential functions for the dimer. Comparisons were made with exptl. thermodn. and structural data including the neutron diffraction results of Thiessen and Narten (1982). The computed densities and potential energies agree with expt. except for the original Bernal-Fowler model, which yields an 18% overest. of the d. and poor structural results. The discrepancy may be due to the correction terms needed in processing the neutron data or to an effect uniformly neglected in the computations. Comparisons were made for the self-diffusion coeffs. obtained from mol. dynamics simulations.
- 72MacKerell, A. D., Jr; Bashford, D.; Bellott, M.; Dunbrack, R. L., Jr; Evanseck, J. D.; Field, M. J.; Fischer, S.; Gao, J.; Guo, H.; Ha, S. All-atom empirical potential for molecular modeling and dynamics studies of proteins. J. Phys. Chem. B 1998, 102, 3586– 3616, DOI: 10.1021/jp973084fGoogle Scholar72https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXivVOlsb4%253D&md5=ebb5100dafd0daeee60ca2fa66c1324aAll-Atom Empirical Potential for Molecular Modeling and Dynamics Studies of ProteinsMacKerell, A. D., Jr.; Bashford, D.; Bellott, M.; Dunbrack, R. L.; Evanseck, J. D.; Field, M. J.; Fischer, S.; Gao, J.; Guo, H.; Ha, S.; Joseph-McCarthy, D.; Kuchnir, L.; Kuczera, K.; Lau, F. T. K.; Mattos, C.; Michnick, S.; Ngo, T.; Nguyen, D. T.; Prodhom, B.; Reiher, W. E., III; Roux, B.; Schlenkrich, M.; Smith, J. C.; Stote, R.; Straub, J.; Watanabe, M.; Wiorkiewicz-Kuczera, J.; Yin, D.; Karplus, M.Journal of Physical Chemistry B (1998), 102 (18), 3586-3616CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)New protein parameters are reported for the all-atom empirical energy function in the CHARMM program. The parameter evaluation was based on a self-consistent approach designed to achieve a balance between the internal (bonding) and interaction (nonbonding) terms of the force field and among the solvent-solvent, solvent-solute, and solute-solute interactions. Optimization of the internal parameters used exptl. gas-phase geometries, vibrational spectra, and torsional energy surfaces supplemented with ab initio results. The peptide backbone bonding parameters were optimized with respect to data for N-methylacetamide and the alanine dipeptide. The interaction parameters, particularly the at. charges, were detd. by fitting ab initio interaction energies and geometries of complexes between water and model compds. that represented the backbone and the various side chains. In addn., dipole moments, exptl. heats and free energies of vaporization, solvation and sublimation, mol. vols., and crystal pressures and structures were used in the optimization. The resulting protein parameters were tested by applying them to noncyclic tripeptide crystals, cyclic peptide crystals, and the proteins crambin, bovine pancreatic trypsin inhibitor, and carbonmonoxy myoglobin in vacuo and in a crystal. A detailed anal. of the relationship between the alanine dipeptide potential energy surface and calcd. protein φ, χ angles was made and used in optimizing the peptide group torsional parameters. The results demonstrate that use of ab initio structural and energetic data by themselves are not sufficient to obtain an adequate backbone representation for peptides and proteins in soln. and in crystals. Extensive comparisons between mol. dynamics simulation and exptl. data for polypeptides and proteins were performed for both structural and dynamic properties. Calcd. data from energy minimization and dynamics simulations for crystals demonstrate that the latter are needed to obtain meaningful comparisons with exptl. crystal structures. The presented parameters, in combination with the previously published CHARMM all-atom parameters for nucleic acids and lipids, provide a consistent set for condensed-phase simulations of a wide variety of mols. of biol. interest.
- 73Venable, R. M.; Luo, Y.; Gawrisch, K.; Roux, B.; Pastor, R. W. Simulations of anionic lipid membranes: development of interaction-specific ion parameters and validation using NMR data. J. Phys. Chem. B 2013, 117, 10183– 10192, DOI: 10.1021/jp401512zGoogle Scholar73https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXht1ars7%252FL&md5=f479595c0149e02b4b9e663b772120b9Simulations of Anionic Lipid Membranes: Development of Interaction-Specific Ion Parameters and Validation Using NMR DataVenable, Richard M.; Luo, Yun; Gawrisch, Klaus; Roux, Benoit; Pastor, Richard W.Journal of Physical Chemistry B (2013), 117 (35), 10183-10192CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)Overbinding of ions to lipid head groups is a potentially serious artifact in simulations of charged lipid bilayers. In this study, the Lennard-Jones radii in the CHARMM force field for interactions of Na+ and lipid oxygen atoms of carboxyl, phosphate, and ester groups were revised to match osmotic pressure data on sodium acetate and electrophoresis data on palmitoyloleoyl phosphatidylcholine (POPC) vesicles. The new parameters were then validated by successfully reproducing previously published exptl. NMR deuterium order parameters for dimyristoyl phosphatidylglycerol (DMPG) and newly obtained values for palmitoyloleoyl phosphatidylserine (POPS). Although the increases in Lennard-Jones diams. are only 0.02-0.12 Å, they are sufficient to reduce Na+ binding, and thereby increase surface areas per lipid by 5-10% compared with the unmodified parameters.
- 74Abraham, M. J.; Murtola, T.; Schulz, R.; Páll, S.; Smith, J. C.; Hess, B.; Lindahl, E. GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 2015, 1-2, 19– 25, DOI: 10.1016/j.softx.2015.06.001Google ScholarThere is no corresponding record for this reference.
- 75Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling through velocity rescaling. J. Chem. Phys. 2007, 126, 014101 DOI: 10.1063/1.2408420Google Scholar75https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXosVCltg%253D%253D&md5=9c182b57bfc65bca6be23c8c76b4be77Canonical sampling through velocity rescalingBussi, Giovanni; Donadio, Davide; Parrinello, MicheleJournal of Chemical Physics (2007), 126 (1), 014101/1-014101/7CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The authors present a new mol. dynamics algorithm for sampling the canonical distribution. In this approach the velocities of all the particles are rescaled by a properly chosen random factor. The algorithm is formally justified and it is shown that, in spite of its stochastic nature, a quantity can still be defined that remains const. during the evolution. In numerical applications this quantity can be used to measure the accuracy of the sampling. The authors illustrate the properties of this new method on Lennard-Jones and TIP4P water models in the solid and liq. phases. Its performance is excellent and largely independent of the thermostat parameter also with regard to the dynamic properties.
- 76Parrinello, M.; Rahman, A. Polymorphic transitions in single crystals: A new molecular dynamics method. J. Appl. Phys. 1981, 52, 7182– 7190, DOI: 10.1063/1.328693Google Scholar76https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL38XislSnuw%253D%253D&md5=a0a5617389f6cabbf2a405c649aadf03Polymorphic transitions in single crystals: A new molecular dynamics methodParrinello, M.; Rahman, A.Journal of Applied Physics (1981), 52 (12), 7182-90CODEN: JAPIAU; ISSN:0021-8979.A Lagrangian formulation is introduced; it can be used to make mol. dynamics (MD) calcns. on systems under the most general, externally applied, conditions of stress. In this formulation the MD cell shape and size can change according to dynamic equations given by this Lagrangian. This MD technique was used to the study of structural transitions of a Ni single crystal under uniform uniaxial compressive and tensile loads. Some results regarding the stress-strain relation obtained by static calcns. are invalid at finite temp. Under compressive loading, the model of Ni shows a bifurcation in its stress-strain relation; this bifurcation provides a link in configuration space between cubic and hexagonal close packing. Such a transition could perhaps be obsd. exptl. under extreme conditions of shock.
- 77Hess, B. P-LINCS: A A parallel linear constraint solver for molecular simulation. J. Chem. Theory Comput. 2008, 4, 116– 122, DOI: 10.1021/ct700200bGoogle Scholar77https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXhtlKru7zL&md5=476d5ca2eb25574d44b775996fff7b75P-LINCS: A Parallel Linear Constraint Solver for Molecular SimulationHess, BerkJournal of Chemical Theory and Computation (2008), 4 (1), 116-122CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)By removing the fastest degrees of freedom, constraints allow for an increase of the time step in mol. simulations. In the last decade parallel simulations have become commonplace. However, up till now efficient parallel constraint algorithms have not been used with domain decompn. In this paper the parallel linear constraint solver (P-LINCS) is presented, which allows the constraining of all bonds in macromols. Addnl. the energy conservation properties of (P-)LINCS are assessed in view of improvements in the accuracy of uncoupled angle constraints and integration in single precision.
- 78Darden, T.; York, D.; Pedersen, L. Particle mesh Ewald: An N · log(N) method for Ewald sums in large systems. J. Chem. Phys. 1993, 98, 10089– 10092, DOI: 10.1063/1.464397Google Scholar78https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXks1Ohsr0%253D&md5=3c9f230bd01b7b714fd096d4d2e755f6Particle mesh Ewald: an N·log(N) method for Ewald sums in large systemsDarden, Tom; York, Darrin; Pedersen, LeeJournal of Chemical Physics (1993), 98 (12), 10089-92CODEN: JCPSA6; ISSN:0021-9606.An N·log(N) method for evaluating electrostatic energies and forces of large periodic systems is presented. The method is based on interpolation of the reciprocal space Ewald sums and evaluation of the resulting convolution using fast Fourier transforms. Timings and accuracies are presented for three large cryst. ionic systems.
- 79Loche, P.; Steinbrunner, P.; Friedowitz, S.; Netz, R. R.; Bonthuis, D. J. Transferable Ion Force Fields in Water from a Simultaneous Optimization of Ion Solvation and Ion–Ion Interaction. J. Phys. Chem. B 2021, 125, 8581– 8587, DOI: 10.1021/acs.jpcb.1c05303Google Scholar79https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhsF2lur3L&md5=f632bce7d53b0fc8f8d8763205915a02Transferable Ion Force Fields in Water from a Simultaneous Optimization of Ion Solvation and Ion-Ion InteractionLoche, Philip; Steinbrunner, Patrick; Friedowitz, Sean; Netz, Roland R.; Bonthuis, Douwe JanJournal of Physical Chemistry B (2021), 125 (30), 8581-8587CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)The poor performance of many existing nonpolarizable ion force fields is typically blamed on either the lack of explicit polarizability, the absence of charge transfer, or the use of unreduced Coulomb interactions. However, this anal. disregards the large and mostly unexplored parameter range offered by the Lennard-Jones potential. We use a global optimization procedure to develop water-model-transferable force fields for the ions K+, Na+, Cl-, and Br- in the complete parameter space of all Lennard-Jones interactions using std. mixing rules. No extra-thermodn. assumption is necessary for the simultaneous optimization of the four ion pairs. After optimization with respect to the exptl. solvation free energy and activity, the force fields reproduce the concn. dependent d., ionic cond. and dielec. const. with high accuracy. The force field is fully transferable between SPC/E, TIP3P, and TIP4P/ε water models. Our results show that a thermodynamically consistent force field for these ions needs only Lennard-Jones and std. Coulomb interactions.
- 80Humphrey, W.; Dalke, A.; Schulten, K. VMD: visual molecular dynamics. J. Mol. Graphics 1996, 14, 33– 38, DOI: 10.1016/0263-7855(96)00018-5Google Scholar80https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28Xis12nsrg%253D&md5=1e3094ec3151fb85c5ff05f8505c78d5VDM: visual molecular dynamicsHumphrey, William; Dalke, Andrew; Schulten, KlausJournal of Molecular Graphics (1996), 14 (1), 33-8, plates, 27-28CODEN: JMGRDV; ISSN:0263-7855. (Elsevier)VMD is a mol. graphics program designed for the display and anal. of mol. assemblies, in particular, biopolymers such as proteins and nucleic acids. VMD can simultaneously display any no. of structures using a wide variety of rendering styles and coloring methods. Mols. are displayed as one or more "representations," in which each representation embodies a particular rendering method and coloring scheme for a selected subset of atoms. The atoms displayed in each representation are chosen using an extensive atom selection syntax, which includes Boolean operators and regular expressions. VMD provides a complete graphical user interface for program control, as well as a text interface using the Tcl embeddable parser to allow for complex scripts with variable substitution, control loops, and function calls. Full session logging is supported, which produces a VMD command script for later playback. High-resoln. raster images of displayed mols. may be produced by generating input scripts for use by a no. of photorealistic image-rendering applications. VMD has also been expressly designed with the ability to animate mol. dynamics (MD) simulation trajectories, imported either from files or from a direct connection to a running MD simulation. VMD is the visualization component of MDScope, a set of tools for interactive problem solving in structural biol., which also includes the parallel MD program NAMD, and the MDCOMM software used to connect the visualization and simulation programs, VMD is written in C++, using an object-oriented design; the program, including source code and extensive documentation, is freely available via anonymous ftp and through the World Wide Web.
- 81Gowers, R. J.; Linke, M.; Barnoud, J.; Reddy, T. J.; Melo, M. N.; Seyler, S. L.; Domanski, J.; Dotson, D. L.; Buchoux, S.; Kenney, I. M.; Beckstein, O. MDAnalysis: A Python Package for the Rapid Analysis of Molecular Dynamics Simulations; Proceedings of the 15th Python in Science Conference; Office of Scientific and Technical Information: Los Alamos, NM, 2016; p 105.Google ScholarThere is no corresponding record for this reference.
- 82Michaud-Agrawal, N.; Denning, E. J.; Woolf, T. B.; Beckstein, O. MDAnalysis: a toolkit for the analysis of molecular dynamics simulations. J. Comput. Chem. 2011, 32, 2319– 2327, DOI: 10.1002/jcc.21787Google Scholar82https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXnvFalsr8%253D&md5=d567042c65cfdc1c81336a29137654bfMDAnalysis: A toolkit for the analysis of molecular dynamics simulationsMichaud-Agrawal, Naveen; Denning, Elizabeth J.; Woolf, Thomas B.; Beckstein, OliverJournal of Computational Chemistry (2011), 32 (10), 2319-2327CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)MDAnal. is an object-oriented library for structural and temporal anal. of mol. dynamics (MD) simulation trajectories and individual protein structures. It is written in the Python language with some performance-crit. code in C. It uses the powerful NumPy package to expose trajectory data as fast and efficient NumPy arrays. It has been tested on systems of millions of particles. Many common file formats of simulation packages including CHARMM, Gromacs, Amber, and NAMD and the Protein Data Bank format can be read and written. Atoms can be selected with a syntax similar to CHARMM's powerful selection commands. MDAnal. enables both novice and experienced programmers to rapidly write their own anal. tools and access data stored in trajectories in an easily accessible manner that facilitates interactive explorative anal. MDAnal. has been tested on and works for most Unix-based platforms such as Linux and Mac OS X. It is freely available under the GNU General Public License from http://mdanal.googlecode.com. © 2011 Wiley Periodicals, Inc. J Comput Chem 2011.
- 83Schwarzl, R.; Liese, S.; Brünig, F. N.; Laudisio, F.; Netz, R. R. Force Response of Polypeptide Chains from Water-Explicit MD Simulations. Macromolecules 2020, 53, 4618– 4629, DOI: 10.1021/acs.macromol.0c00138Google Scholar83https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhtFeltrnK&md5=0592929f6ff286df1de443f5f0c3748fForce Response of Polypeptide Chains from Water-Explicit MD SimulationsSchwarzl, Richard; Liese, Susanne; Bruenig, Florian N.; Laudisio, Fabio; Netz, Roland R.Macromolecules (Washington, DC, United States) (2020), 53 (12), 4618-4629CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)Using mol. dynamics simulations in explicit water, the force-extension relations for the five homopeptides polyglycine, polyalanine, polyasparagine, poly(glutamic acid), and polylysine are investigated. From simulations in the low-force regime the Kuhn length is detd., from simulations in the high-force regime the equil. contour length and the linear and nonlinear stretching moduli, which agree well with quantum-chem. d.-functional theory calcns., are detd. All these parameters vary considerably between the different polypeptides. The augmented inhomogeneous partially freely rotating chain (iPFRC) model, which accounts for side-chain interactions and restricted dihedral rotation, is demonstrated to describe the simulated force-extension relations very well. We present a quant. comparison between published exptl. single-mol. force-extension curves for different polypeptides with simulation and model predictions. The thermodn. stretching properties of polypeptides are investigated by decompn. of the stretching free energy into energetic and entropic contributions.
- 84Flyvbjerg, H.; Petersen, H. G. Error estimates on averages of correlated data. J. Chem. Phys. 1989, 91, 461– 466, DOI: 10.1063/1.457480Google Scholar84https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXkslOis7s%253D&md5=a424bd42502aa2989d102aa4c87b8ecaError estimates on averages of correlated dataFlyvbjerg, H.; Petersen, H. G.Journal of Chemical Physics (1989), 91 (1), 461-6CODEN: JCPSA6; ISSN:0021-9606.A description is given on how the true statistical error on an av. of correlated data can be obtained with ease and efficiency by a renormalization group method. The method is illustrated with numerical and anal. examples, having finite as well as infinite range correlations.
Cited By
This article has not yet been cited by other publications.
Article Views
Altmetric
Citations
Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.
Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.
The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.
Recommended Articles
References
This article references 84 other publications.
- 1Kornberg, R. D.; Lorch, Y. Twenty-five years of the nucleosome, fundamental particle of the eukaryote chromosome. Cell 1999, 98, 285– 294, DOI: 10.1016/S0092-8674(00)81958-31https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXlt1Cgt7o%253D&md5=7996cade955faa759d9bca55759eb513Twenty-five years of the nucleosome, fundamental particle of the eukaryote chromosomeKornberg, Roger D.; Lorch, YahliCell (Cambridge, Massachusetts) (1999), 98 (3), 285-294CODEN: CELLB5; ISSN:0092-8674. (Cell Press)A review with ∼100 refs., including sections entitled the histones, the nucleosome, structure of the nucleosome, structure of chromatin fibers, nucleosomes repress transcription, histone acetylation, histone deacetylation, chromatin remodeling silencing, nucleosomes and transcription elongation, multiple mechanisms of activation and repression, viral infection and cancer, general principles, and perspectives.
- 2Kunze, K.-K.; Netz, R. Complexes of semiflexible polyelectrolytes and charged spheres as models for salt-modulated nucleosomal structures. Phys. Rev. E 2002, 66, 011918 DOI: 10.1103/PhysRevE.66.011918There is no corresponding record for this reference.
- 3Parsaeian, A.; De La Cruz, M. O.; Marko, J. F. Binding-rebinding dynamics of proteins interacting nonspecifically with a long DNA molecule. Phys. Rev. E 2013, 88, 040703 DOI: 10.1103/PhysRevE.88.040703There is no corresponding record for this reference.
- 4Shin, Y.; Brangwynne, C. P. Liquid phase condensation in cell physiology and disease. Science 2017, 357, eaaf4382 DOI: 10.1126/science.aaf4382There is no corresponding record for this reference.
- 5Goedert, M.; Jakes, R.; Spillantini, M.; Hasegawa, M.; Smith, M.; Crowther, R. Assembly of microtubule-associated protein tau into Alzheimer-like filaments induced by sulphated glycosaminoglycans. Nature 1996, 383, 550– 553, DOI: 10.1038/383550a05https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28Xmt1GqsL8%253D&md5=8c572e7db12bc47fa3fa8be625421fa8Assembly of microtubule-associated protein tau into Alzheimer-like filaments induced by sulfated glycosaminoglycansGoedert, M.; Jakes, R.; Spillantini, M. G.; Hasegawa, M.; Smith, M. J.; Crowther, R. A.Nature (London) (1996), 383 (6600), 550-553CODEN: NATUAS; ISSN:0028-0836. (Macmillan Magazines)The paired helical filament (PHF) is the major component of the neurofibrillary deposits that form a defining neuropathol. characteristic of Alzheimer's disease. PHFs are composed of microtubule-assocd. protein tau, in a hyperphosphorylated state. Hyperphosphorylation of tau results in its inability to bind to microtubules and is believed to precede PHF assembly. However, it is unclear whether hyperphosphorylation of tau is either necessary or sufficient for PHF formation. Here we show that non-phosphorylated recombinant tau isoforms with three microtubule-binding repeats form paired helical-like filaments under physiol. conditions in vitro, when incubated with sulfated glycosaminoglycans such as heparin or heparan sulfate. Furthermore, heparin prevents tau from binding to microtubules and promotes microtubule disassembly. Finally, we show that heparan sulfate and hyperphosphorylated tau coexist in nerve cells of the Alzheimer's disease brain at the earliest known stages of neurofibrillary pathol. These findings, with previous studies which show that heparin stimulates tau phosphorylation by a no. of protein kinases, indicate that sulfated glycosaminoglycans may be a key factor in the formation of the neurofibrillary lesions of Alzheimer's disease.
- 6Fichou, Y.; Lin, Y.; Rauch, J. N.; Vigers, M.; Zeng, Z.; Srivastava, M.; Keller, T. J.; Freed, J. H.; Kosik, K. S.; Han, S. Cofactors are essential constituents of stable and seeding-active tau fibrils. Proc. Natl. Acad. Sci. U.S.A. 2018, 115, 13234– 13239, DOI: 10.1073/pnas.18100581156https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXisF2rsrvL&md5=47c4d658171711202a8be7cbddf4535fCofactors are essential constituents of stable and seeding-active tau fibrilsFichou, Yann; Lin, Yanxian; Rauch, Jennifer N.; Vigers, Michael; Zeng, Zhikai; Srivastava, Madhur; Keller, Timothy J.; Freed, Jack H.; Kosik, Kenneth S.; Han, SongiProceedings of the National Academy of Sciences of the United States of America (2018), 115 (52), 13234-13239CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)Amyloid fibrils are cross-β-rich aggregates that are exceptionally stable forms of protein assembly. Accumulation of tau amyloid fibrils is involved in many neurodegenerative diseases, including Alzheimer's disease (AD). Heparin-induced aggregates have been widely used and assumed to be a good tau amyloid fibril model for most biophys. studies. Here we show that mature fibrils made of 4R tau variants, prepd. with heparin or RNA, spontaneously depolymerize and release monomers when their cofactors are removed. We demonstrate that the cross-β-sheet assembly formed in vitro with polyanion addn. is unstable at room temp. We furthermore demonstrate high seeding capacity with transgenic AD mouse brain-extd. tau fibrils in vitro that, however, is exhausted after one generation, while supplementation with RNA cofactors resulted in sustained seeding over multiple generations. We suggest that tau fibrils formed in brains are supported by unknown cofactors and inhere higher-quality packing, as reflected in a more distinct conformational arrangement in the mouse fibril-seeded, compared with heparin-induced, tau fibrils. Our study suggests that the role of cofactors in tauopathies is a worthy focus of future studies, as they may be viable targets for diagnosis and therapeutics.
- 7Achazi, K.; Haag, R.; Ballauff, M.; Dernedde, J.; Kizhakkedathu, J. N.; Maysinger, D.; Multhaup, G. Understanding the interaction of polyelectrolyte architectures with proteins and biosystems. Angew. Chem., Int. Ed. 2021, 60, 3882– 3904, DOI: 10.1002/anie.2020064577https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXitFyhs7nM&md5=84aa4dbce05059373ed72fc46bd469cbUnderstanding the Interaction of Polyelectrolyte Architectures with Proteins and BiosystemsAchazi, Katharina; Haag, Rainer; Ballauff, Matthias; Dernedde, Jens; Kizhakkedathu, Jayachandran N.; Maysinger, Dusica; Multhaup, GerdAngewandte Chemie, International Edition (2021), 60 (8), 3882-3904CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. The counterions neutralizing the charges on polyelectrolytes such as DNA or heparin may dissoc. in water and greatly influence the interaction of such polyelectrolytes with biomols., particularly proteins. In this Review the authors give an overview of studies on the interaction of proteins with polyelectrolytes and how this knowledge can be used for medical applications. Counterion release was identified as the main driving force for the binding of proteins to polyelectrolytes: Patches of pos. charge become multivalent counterions of the polyelectrolyte and lead to the release of counterions from the polyelectrolyte and a concomitant increase in entropy. This is shown from studies on the interaction of proteins with natural and synthetic polyelectrolytes. Special emphasis is paid to sulfated dendritic polyglycerols (dPGS). The Review demonstrates that the authors are moving to a better understanding of charge-charge interactions in systems of biol. relevance. Research along these lines will aid and promote the design of synthetic polyelectrolytes for medical applications.
- 8Kayitmazer, A. B.; Seeman, D.; Minsky, B. B.; Dubin, P. L.; Xu, Y. Protein–polyelectrolyte interactions. Soft Matter 2013, 9, 2553– 2583, DOI: 10.1039/c2sm27002a8https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXitVCitr0%253D&md5=652b0fe09debf23e810f733326abab58Protein-polyelectrolyte interactionsKayitmazer, A. Basak; Seeman, Daniel; Minsky, Burcu Baykal; Dubin, Paul L.; Xu, YishengSoft Matter (2013), 9 (9), 2553-2583CODEN: SMOABF; ISSN:1744-683X. (Royal Society of Chemistry)A review. The interactions of proteins and polyelectrolytes lead to diverse applications in sepns., delivery and wound repair, and are thus of interest to scientists in e.g. (a) glycobiol., (b) tissue engineering, (c) biosensing, and (d) pharmacol. This breadth is accompanied by an assortment of contexts and models in which polyelectrolytes are seen as (a) protein cognates assisting in complex cellular roles, (b) surrogates for the extracellular matrix, mimicking its hydration, mech. and sequestering properties, (c) benign hosts that gently entrap, deposit and tether protein substrate specificity, and (d) selective but non-specific agents that modify protein soly. Unsurprisingly, this literature is somewhat segregated by objectives and paradigms. We hope this review, which emphasizes publications over the last 8 years, represents and also counterbalances that divergence. An ongoing theme is the role of electrostatics, and we show how this leads to the variety of phys. forms taken by protein-polyelectrolyte complexes. We present approaches towards anal. and characterization, motivated by the goal of structure-property elucidation. Such understanding should guide in applications, our third topic. We present recent developments in modeling and simulations of protein-polyelectrolyte systems. We close with a prospective on future developments in this field.
- 9Nandy, B.; Saurabh, S.; Sahoo, A. K.; Dixit, N. M.; Maiti, P. K. The SPL7013 dendrimer destabilizes the HIV-1 gp120-CD4 complex. Nanoscale 2015, 7, 18628– 18641, DOI: 10.1039/C5NR04632G9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhs1Squ7bF&md5=3a1bef77f1f9ac6af6f04778be48045eThe SPL7013 dendrimer destabilizes the HIV-1 gp120-CD4 complexNandy, Bidisha; Saurabh, Suman; Sahoo, Anil Kumar; Dixit, Narendra M.; Maiti, Prabal K.Nanoscale (2015), 7 (44), 18628-18641CODEN: NANOHL; ISSN:2040-3372. (Royal Society of Chemistry)The poly (l-lysine)-based SPL7013 dendrimer with naphthalene disulfonate surface groups blocks the entry of HIV-1 into target cells and is in clin. trials for development as a topical microbicide. Its mechanism of action against R5 HIV-1, the HIV-1 variant implicated in transmission across individuals, remains poorly understood. Using docking and fully atomistic MD simulations, we find that SPL7013 binds tightly to R5 gp120 in the gp120-CD4 complex but weakly to gp120 alone. Further, the binding, although to multiple regions of gp120, does not occlude the CD4 binding site on gp120, suggesting that SPL7013 does not prevent the binding of R5 gp120 to CD4. Using MD simulations to compute binding energies of several docked structures, we find that SPL7013 binding to gp120 significantly weakens the gp120-CD4 complex. Finally, we use steered mol. dynamics (SMD) to study the kinetics of the dissocn. of the gp120-CD4 complex in the absence of the dendrimer and with the dendrimer bound in each of the several stable configurations to gp120. We find that SPL7013 significantly lowers the force required to rupture the gp120-CD4 complex and accelerates its dissocn. Taken together, our findings suggest that SPL7013 compromises the stability of the R5 gp120-CD4 complex, potentially preventing the accrual of the requisite no. of gp120-CD4 complexes across the virus-cell interface, thereby blocking virus entry.
- 10Cagno, V.; Tseligka, E. D.; Jones, S. T.; Tapparel, C. Heparan sulfate proteoglycans and viral attachment: true receptors or adaptation bias?. Viruses 2019, 11, 596 DOI: 10.3390/v1107059610https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXisVOgu7jO&md5=3cf85bf40431c53df1a530498ddabf4aHeparan sulfate proteoglycans and viral attachment: true receptors or adaptation bias?Cagno, Valeria; Tseligka, Eirini D.; Jones, Samuel T.; Tapparel, CarolineViruses (2019), 11 (7), 596CODEN: VIRUBR; ISSN:1999-4915. (MDPI AG)A review. Heparan sulfate proteoglycans (HSPG) are composed of unbranched, neg. charged heparan sulfate (HS) polysaccharides attached to a variety of cell surface or extracellular matrix proteins. Widely expressed, they mediate many biol. activities, including angiogenesis, blood coagulation, developmental processes, and cell homeostasis. HSPG are highly sulfated and broadly used by a range of pathogens, esp. viruses, to attach to the cell surface. In this review, we summarize the current knowledge on HSPG-virus interactions and distinguish viruses with established HS binding, viruses that bind HS only after intra-host or cell culture adaptation, and finally, viruses whose dependence on HS for infection is debated. We also provide an overview of the antiviral compds. designed to interfere with HS binding. Many questions remain about the true importance of these receptors in vivo, knowledge that is crit. for the design of future antiviral therapies.
- 11Liese, S.; Netz, R. R. Quantitative prediction of multivalent ligand-receptor binding affinities for influenza, cholera, and anthrax inhibition. ACS Nano 2018, 12, 4140– 4147, DOI: 10.1021/acsnano.7b0847911https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXjtlOnsr8%253D&md5=8a1df12e15cf156042bec1a0cadc51daQuantitative Prediction of Multivalent Ligand-Receptor Binding Affinities for Influenza, Cholera, and Anthrax InhibitionLiese, Susanne; Netz, Roland R.ACS Nano (2018), 12 (5), 4140-4147CODEN: ANCAC3; ISSN:1936-0851. (American Chemical Society)Multivalency achieves strong, yet reversible binding by the simultaneous formation of multiple weak bonds. It is a key interaction principle in biol. and promising for the synthesis of high-affinity inhibitors of pathogens. We present a mol. model for the binding affinity of synthetic multivalent ligands onto multivalent receptors consisting of n receptor units arranged on a regular polygon. Ligands consist of a geometrically matching rigid polygonal core to which monovalent ligand units are attached via flexible linker polymers, closely mimicking existing exptl. designs. The calcd. binding affinities quant. agree with exptl. studies for cholera toxin (n = 5) and anthrax receptor (n = 7) and allow to predict optimal core size and optimal linker length. Maximal binding affinity is achieved for a core that matches the receptor size and for linkers that have an equil. end-to-end distance that is slightly longer than the geometric sepn. between ligand core and receptor sites. Linkers that are longer than optimal are greatly preferable compared to shorter linkers. The angular steric restriction between ligand unit and linker polymer is shown to be a key parameter. We construct an enhancement diagram that quantifies the multivalent binding affinity compared to monovalent ligands. We conclude that multivalent ligands against influenza viral hemagglutinin (n = 3), cholera toxin (n = 5), and anthrax receptor (n = 7) can outperform monovalent ligands only for a monovalent ligand affinity that exceeds a core-size dependent threshold value. Thus, multivalent drug design needs to balance core size, linker length, as well as monovalent ligand unit affinity.
- 12Lauster, D.; Osterrieder, K.; Haag, R.; Ballauff, M.; Herrmann, A. Respiratory viruses interacting with cells: The importance of electrostatics. Front. Microbiol. 2023, 14, 1169547 DOI: 10.3389/fmicb.2023.1169547There is no corresponding record for this reference.
- 13Chu, H.; Hu, B.; Huang, X.; Chai, Y.; Zhou, D.; Wang, Y.; Shuai, H.; Yang, D.; Hou, Y.; Zhang, X. Host and viral determinants for efficient SARS-CoV-2 infection of the human lung. Nat. Commun. 2021, 12, 134 DOI: 10.1038/s41467-020-20457-w13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhtF2gurg%253D&md5=637c93ff402b13e3a45c2e5dc53548d7Host and viral determinants for efficient SARS-CoV-2 infection of the human lungChu, Hin; Hu, Bingjie; Huang, Xiner; Chai, Yue; Zhou, Dongyan; Wang, Yixin; Shuai, Huiping; Yang, Dong; Hou, Yuxin; Zhang, Xi; Yuen, Terrence Tsz-Tai; Cai, Jian-Piao; Zhang, Anna Jinxia; Zhou, Jie; Yuan, Shuofeng; To, Kelvin Kai-Wang; Chan, Ivy Hau-Yee; Sit, Ko-Yung; Foo, Dominic Chi-Chung; Wong, Ian Yu-Hong; Ng, Ada Tsui-Lin; Cheung, Tan To; Law, Simon Ying-Kit; Au, Wing-Kuk; Brindley, Melinda A.; Chen, Zhiwei; Kok, Kin-Hang; Chan, Jasper Fuk-Woo; Yuen, Kwok-YungNature Communications (2021), 12 (1), 134CODEN: NCAOBW; ISSN:2041-1723. (Nature Research)Abstr.: Understanding the factors that contribute to efficient SARS-CoV-2 infection of human cells may provide insights on SARS-CoV-2 transmissibility and pathogenesis, and reveal targets of intervention. Here, we analyze host and viral determinants essential for efficient SARS-CoV-2 infection in both human lung epithelial cells and ex vivo human lung tissues. We identify heparan sulfate as an important attachment factor for SARS-CoV-2 infection. Next, we show that sialic acids present on ACE2 prevent efficient spike/ACE2-interaction. While SARS-CoV infection is substantially limited by the sialic acid-mediated restriction in both human lung epithelial cells and ex vivo human lung tissues, infection by SARS-CoV-2 is limited to a lesser extent. We further demonstrate that the furin-like cleavage site in SARS-CoV-2 spike is required for efficient virus replication in human lung but not intestinal tissues. These findings provide insights on the efficient SARS-CoV-2 infection of human lungs.
- 14Kim, S. H.; Kearns, F. L.; Rosenfeld, M. A.; Votapka, L.; Casalino, L.; Papanikolas, M.; Amaro, R. E.; Freeman, R. SARS-CoV-2 evolved variants optimize binding to cellular glycocalyx. Cell Rep. Phys. Sci. 2023, 4, 101346 DOI: 10.1016/j.xcrp.2023.101346There is no corresponding record for this reference.
- 15Nie, C.; Pouyan, P.; Lauster, D.; Trimpert, J.; Kerkhoff, Y.; Szekeres, G. P.; Wallert, M.; Block, S.; Sahoo, A. K.; Dernedde, J.; Pagel, K.; Kaufer, B. B.; Netz, R. R.; Ballauff, M.; Haag, R. Polysulfates block SARS-CoV-2 uptake via electrostatic interactions. Angew. Chem., Int. Ed. 2021, 60, 15870– 15878, DOI: 10.1002/anie.20210271715https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhsVKmtb7I&md5=681dca8bfea352c705f4ece7f110c6e3Polysulfates Block SARS-CoV-2 Uptake through Electrostatic InteractionsNie, Chuanxiong; Pouyan, Paria; Lauster, Daniel; Trimpert, Jakob; Kerkhoff, Yannic; Szekeres, Gergo Peter; Wallert, Matthias; Block, Stephan; Sahoo, Anil Kumar; Dernedde, Jens; Pagel, Kevin; Kaufer, Benedikt B.; Netz, Roland R.; Ballauff, Matthias; Haag, RainerAngewandte Chemie, International Edition (2021), 60 (29), 15870-15878CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)Here we report that neg. charged polysulfates can bind to the spike protein of SARS-CoV-2 via electrostatic interactions. Using a plaque redn. assay, we compare inhibition of SARS-CoV-2 by heparin, pentosan sulfate, linear polyglycerol sulfate (LPGS) and hyperbranched polyglycerol sulfate (HPGS). Highly sulfated LPGS is the optimal inhibitor, with an IC50 of 67μg mL-1 (approx. 1.6μM). This synthetic polysulfate exhibits more than 60-fold higher virus inhibitory activity than heparin (IC50: 4084μg mL-1), along with much lower anticoagulant activity. Furthermore, in mol. dynamics simulations, we verified that LPGS can bind more strongly to the spike protein than heparin, and that LPGS can interact even more with the spike protein of the new N501Y and E484K variants. Our study demonstrates that the entry of SARS-CoV-2 into host cells can be blocked via electrostatic interactions, therefore LPGS can serve as a blueprint for the design of novel viral inhibitors of SARS-CoV-2.
- 16Clausen, T. M.; Sandoval, D. R.; Spliid, C. B.; Pihl, J.; Perrett, H. R.; Painter, C. D.; Narayanan, A.; Majowicz, S. A.; Kwong, E. M.; McVicar, R. N. SARS-CoV-2 infection depends on cellular heparan sulfate and ACE2. Cell 2020, 183, 1043– 1057, DOI: 10.1016/j.cell.2020.09.03316https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhvFGqsLjK&md5=f478b8b320b347ca7f8accca25b822bcSARS-CoV-2 Infection Depends on Cellular Heparan Sulfate and ACE2Clausen, Thomas Mandel; Sandoval, Daniel R.; Spliid, Charlotte B.; Pihl, Jessica; Perrett, Hailee R.; Painter, Chelsea D.; Narayanan, Anoop; Majowicz, Sydney A.; Kwong, Elizabeth M.; McVicar, Rachael N.; Thacker, Bryan E.; Glass, Charles A.; Yang, Zhang; Torres, Jonathan L.; Golden, Gregory J.; Bartels, Phillip L.; Porell, Ryan N.; Garretson, Aaron F.; Laubach, Logan; Feldman, Jared; Yin, Xin; Pu, Yuan; Hauser, Blake M.; Caradonna, Timothy M.; Kellman, Benjamin P.; Martino, Cameron; Gordts, Philip L. S. M.; Chanda, Sumit K.; Schmidt, Aaron G.; Godula, Kamil; Leibel, Sandra L.; Jose, Joyce; Corbett, Kevin D.; Ward, Andrew B.; Carlin, Aaron F.; Esko, Jeffrey D.Cell (Cambridge, MA, United States) (2020), 183 (4), 1043-1057.e15CODEN: CELLB5; ISSN:0092-8674. (Cell Press)We show that SARS-CoV-2 spike protein interacts with both cellular heparan sulfate and angiotensin-converting enzyme 2 (ACE2) through its receptor-binding domain (RBD). Docking studies suggest a heparin/heparan sulfate-binding site adjacent to the ACE2-binding site. Both ACE2 and heparin can bind independently to spike protein in vitro, and a ternary complex can be generated using heparin as a scaffold. Electron micrographs of spike protein suggests that heparin enhances the open conformation of the RBD that binds ACE2. On cells, spike protein binding depends on both heparan sulfate and ACE2. Unfractionated heparin, non-anticoagulant heparin, heparin lyases, and lung heparan sulfate potently block spike protein binding and/or infection by pseudotyped virus and authentic SARS-CoV-2 virus. We suggest a model in which viral attachment and infection involves heparan sulfate-dependent enhancement of binding to ACE2. Manipulation of heparan sulfate or inhibition of viral adhesion by exogenous heparin presents new therapeutic opportunities.
- 17Lever, R.; Page, C. P. Non-Anticoagulant Effects of Heparin: An Overview. In Heparin - A Century of Progress; Lever, R.; Mulloy, B.; Page, C. P., Eds.; Springer: Berlin, Heidelberg, 2012; pp 281– 305.There is no corresponding record for this reference.
- 18Page, C. Heparin and related drugs: beyond anticoagulant activity. ISRN Pharmacol. 2013, 1, 910743 DOI: 10.1155/2013/910743There is no corresponding record for this reference.
- 19Oates, J. A.; Wood, A. J. J.; Hirsh, J. Heparin. N. Engl. J. Med. 1991, 324, 1565– 1574, DOI: 10.1056/NEJM199105303242206There is no corresponding record for this reference.
- 20Cate, H. T. Surviving Covid-19 with Heparin?. N. Engl. J. Med. 2021, 385, 845– 846, DOI: 10.1056/NEJMe2111151There is no corresponding record for this reference.
- 21Lan, J.; Ge, J.; Yu, J.; Shan, S.; Zhou, H.; Fan, S.; Zhang, Q.; Shi, X.; Wang, Q.; Zhang, L.; Wang, X. Structure of the SARS-CoV-2 spike receptor-binding domain bound to the ACE2 receptor. Nature 2020, 581, 215– 220, DOI: 10.1038/s41586-020-2180-521https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXoslOqtL8%253D&md5=279c60143e8e5eb505457e0778baa8efStructure of the SARS-CoV-2 spike receptor-binding domain bound to the ACE2 receptorLan, Jun; Ge, Jiwan; Yu, Jinfang; Shan, Sisi; Zhou, Huan; Fan, Shilong; Zhang, Qi; Shi, Xuanling; Wang, Qisheng; Zhang, Linqi; Wang, XinquanNature (London, United Kingdom) (2020), 581 (7807), 215-220CODEN: NATUAS; ISSN:0028-0836. (Nature Research)Abstr.: A new and highly pathogenic coronavirus (severe acute respiratory syndrome coronavirus-2, SARS-CoV-2) caused an outbreak in Wuhan city, Hubei province, China, starting from Dec. 2019 that quickly spread nationwide and to other countries around the world1-3. Here, to better understand the initial step of infection at an at. level, we detd. the crystal structure of the receptor-binding domain (RBD) of the spike protein of SARS-CoV-2 bound to the cell receptor ACE2. The overall ACE2-binding mode of the SARS-CoV-2 RBD is nearly identical to that of the SARS-CoV RBD, which also uses ACE2 as the cell receptor4. Structural anal. identified residues in the SARS-CoV-2 RBD that are essential for ACE2 binding, the majority of which either are highly conserved or share similar side chain properties with those in the SARS-CoV RBD. Such similarity in structure and sequence strongly indicate convergent evolution between the SARS-CoV-2 and SARS-CoV RBDs for improved binding to ACE2, although SARS-CoV-2 does not cluster within SARS and SARS-related coronaviruses1-3,5. The epitopes of two SARS-CoV antibodies that target the RBD are also analyzed for binding to the SARS-CoV-2 RBD, providing insights into the future identification of cross-reactive antibodies.
- 22Pishko, A. M.; Lefler, D. S.; Gimotty, P.; Paydary, K.; Fardin, S.; Arepally, G. M.; Crowther, M.; Rice, L.; Vega, R.; Cines, D. B. The risk of major bleeding in patients with suspected heparin-induced thrombocytopenia. J. Thromb. Haemostasis 2019, 17, 1956– 1965, DOI: 10.1111/jth.14587There is no corresponding record for this reference.
- 23Nie, C.; Sahoo, A. K.; Netz, R. R.; Herrmann, A.; Ballauff, M.; Haag, R. Charge matters: Mutations in omicron variant favor binding to cells. ChemBioChem 2022, 23, e202100681 DOI: 10.1002/cbic.20210068123https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB38Xislenu7w%253D&md5=69b8f1e9b595a12dbb6ed73d09445a2dCharge Matters: Mutations in Omicron Variant Favor Binding to CellsNie, Chuanxiong; Sahoo, Anil Kumar; Netz, Roland R.; Herrmann, Andreas; Ballauff, Matthias; Haag, RainerChemBioChem (2022), 23 (6), e202100681CODEN: CBCHFX; ISSN:1439-4227. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Evidence is strengthening to suggest that the novel SARS-CoV-2 mutant Omicron, with its more than 60 mutations, will spread and dominate worldwide. Although the mutations in the spike protein are known, the mol. basis for why the addnl. mutations in the spike protein that have not previously occurred account for Omicron's higher infection potential, is not understood. We propose, based on chem. rational and mol. dynamics simulations, that the elevated occurrence of pos. charged amino acids in certain domains of the spike protein (Delta: +4; Omicron: +5 vs. wild type) increases binding to cellular polyanionic receptors, such as heparan sulfate due to multivalent charge-charge interactions. This observation is a starting point for targeted drug development.
- 24Wang, L.; Wu, Y.; Deng, Y.; Kim, B.; Pierce, L.; Krilov, G.; Lupyan, D.; Robinson, S.; Dahlgren, M. K.; Greenwood, J. Accurate and reliable prediction of relative ligand binding potency in prospective drug discovery by way of a modern free-energy calculation protocol and force field. J. Am. Chem. Soc. 2015, 137, 2695– 2703, DOI: 10.1021/ja512751q24https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhsF2iuro%253D&md5=37a4f4a6c085f47ed531342643b6c33bAccurate and Reliable Prediction of Relative Ligand Binding Potency in Prospective Drug Discovery by Way of a Modern Free-Energy Calculation Protocol and Force FieldWang, Lingle; Wu, Yujie; Deng, Yuqing; Kim, Byungchan; Pierce, Levi; Krilov, Goran; Lupyan, Dmitry; Robinson, Shaughnessy; Dahlgren, Markus K.; Greenwood, Jeremy; Romero, Donna L.; Masse, Craig; Knight, Jennifer L.; Steinbrecher, Thomas; Beuming, Thijs; Damm, Wolfgang; Harder, Ed; Sherman, Woody; Brewer, Mark; Wester, Ron; Murcko, Mark; Frye, Leah; Farid, Ramy; Lin, Teng; Mobley, David L.; Jorgensen, William L.; Berne, Bruce J.; Friesner, Richard A.; Abel, RobertJournal of the American Chemical Society (2015), 137 (7), 2695-2703CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Designing tight-binding ligands is a primary objective of small-mol. drug discovery. Over the past few decades, free-energy calcns. have benefited from improved force fields and sampling algorithms, as well as the advent of low-cost parallel computing. However, it has proven to be challenging to reliably achieve the level of accuracy that would be needed to guide lead optimization (∼5× in binding affinity) for a wide range of ligands and protein targets. Not surprisingly, widespread com. application of free-energy simulations has been limited due to the lack of large-scale validation coupled with the tech. challenges traditionally assocd. with running these types of calcns. Here, we report an approach that achieves an unprecedented level of accuracy across a broad range of target classes and ligands, with retrospective results encompassing 200 ligands and a wide variety of chem. perturbations, many of which involve significant changes in ligand chem. structures. In addn., we have applied the method in prospective drug discovery projects and found a significant improvement in the quality of the compds. synthesized that have been predicted to be potent. Compds. predicted to be potent by this approach have a substantial redn. in false positives relative to compds. synthesized on the basis of other computational or medicinal chem. approaches. Furthermore, the results are consistent with those obtained from our retrospective studies, demonstrating the robustness and broad range of applicability of this approach, which can be used to drive decisions in lead optimization.
- 25Cournia, Z.; Allen, B.; Sherman, W. Relative binding free energy calculations in drug discovery: recent advances and practical considerations. J. Chem. Inf. Model. 2017, 57, 2911– 2937, DOI: 10.1021/acs.jcim.7b0056425https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhvFOhu7fM&md5=ffcc9f970ab660eced8d3343ccefeec6Relative Binding Free Energy Calculations in Drug Discovery: Recent Advances and Practical ConsiderationsCournia, Zoe; Allen, Bryce; Sherman, WoodyJournal of Chemical Information and Modeling (2017), 57 (12), 2911-2937CODEN: JCISD8; ISSN:1549-9596. (American Chemical Society)A review. Accurate in silico prediction of protein-ligand binding affinities has been a primary objective of structure-based drug design for decades due to the putative value it would bring to the drug discovery process. However, computational methods have historically failed to deliver value in real-world drug discovery applications due to a variety of scientific, tech., and practical challenges. Recently, a family of approaches commonly referred to as relative binding free energy (RBFE) calcns., which rely on physics-based mol. simulations and statistical mechanics, have shown promise in reliably generating accurate predictions in the context of drug discovery projects. This advance arises from accumulating developments in the underlying scientific methods (decades of research on force fields and sampling algorithms) coupled with vast increases in computational resources (graphics processing units and cloud infrastructures). Mounting evidence from retrospective validation studies, blind challenge predictions, and prospective applications suggests that RBFE simulations can now predict the affinity differences for congeneric ligands with sufficient accuracy and throughput to deliver considerable value in hit-to-lead and lead optimization efforts. Here, the authors present an overview of current RBFE implementations, highlighting recent advances and remaining challenges, along with examples that emphasize practical considerations for obtaining reliable RBFE results. The authors focus specifically on relative binding free energies because the calcns. are less computationally intensive than abs. binding free energy (ABFE) calcns. and map directly onto the hit-to-lead and lead optimization processes, where the prediction of relative binding energies between a ref. mol. and new ideas (virtual mols.) can be used to prioritize mols. for synthesis. The authors describe the crit. aspects of running RBFE calcns., from both theor. and applied perspectives, using a combination of retrospective literature examples and prospective studies from drug discovery projects. This work is intended to provide a contemporary overview of the scientific, tech., and practical issues assocd. with running relative binding free energy simulations, with a focus on real-world drug discovery applications. The authors offer guidelines for improving the accuracy of RBFE simulations, esp. for challenging cases, and emphasize unresolved issues that could be improved by further research in the field.
- 26Mobley, D. L.; Gilson, M. K. Predicting binding free energies: frontiers and benchmarks. Annu. Rev. Biophys. 2017, 46, 531– 558, DOI: 10.1146/annurev-biophys-070816-03365426https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXlvFClsr0%253D&md5=38af2949ba205daaf9899b87c4e1a2bdPredicting Binding Free Energies: Frontiers and BenchmarksMobley, David L.; Gilson, Michael K.Annual Review of Biophysics (2017), 46 (), 531-558CODEN: ARBNCV; ISSN:1936-122X. (Annual Reviews)Binding free energy calcns. based on mol. simulations provide predicted affinities for biomol. complexes. These calcns. begin with a detailed description of a system, including its chem. compn. and the interactions among its components. Simulations of the system are then used to compute thermodn. information, such as binding affinities. Because of their promise for guiding mol. design, these calcns. have recently begun to see widespread applications in early-stage drug discovery. However, many hurdles remain in making them a robust and reliable tool. In this review, we highlight key challenges of these calcns., discuss some examples of these challenges, and call for the designation of std. community benchmark test systems that will help the research community generate and evaluate progress. In our view, progress will require careful assessment and evaluation of new methods, force fields, and modeling innovations on well-characterized benchmark systems, and we lay out our vision for how this can be achieved.
- 27Zhao, J.; Cao, Y.; Zhang, L. Exploring the computational methods for protein-ligand binding site prediction. Comput. Struct. Biotechnol. J. 2020, 18, 417– 426, DOI: 10.1016/j.csbj.2020.02.00827https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXktFeitrY%253D&md5=099de1ae131781598f5d09d48dd23061Exploring the computational methods for protein-ligand binding site predictionZhao, Jingtian; Cao, Yang; Zhang, LeComputational and Structural Biotechnology Journal (2020), 18 (), 417-426CODEN: CSBJAC; ISSN:2001-0370. (Elsevier B.V.)A review. Proteins participate in various essential processes in vivo via interactions with other mols. Identifying the residues participating in these interactions not only provides biol. insights for protein function studies but also has great significance for drug discoveries. Therefore, predicting protein-ligand binding sites has long been under intense research in the fields of bioinformatics and computer aided drug discovery. In this review, we first introduce the research background of predicting protein-ligand binding sites and then classify the methods into four categories, namely, 3D structure-based, template similarity-based, traditional machine learning-based and deep learning-based methods. We describe representative algorithms in each category and elaborate on machine learning and deep learning-based prediction methods in more detail. Finally, we discuss the trends and challenges of the current research such as mol. dynamics simulation based cryptic binding sites prediction, and highlight prospective directions for the near future.
- 28Gapsys, V.; Yildirim, A.; Aldeghi, M.; Khalak, Y.; Van der Spoel, D.; de Groot, B. L. Accurate absolute free energies for ligand–protein binding based on non-equilibrium approaches. Commun. Chem. 2021, 4, 61 DOI: 10.1038/s42004-021-00498-y28https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhvFSktbzO&md5=212f9b8a8ca965e71ec680daa550183fAccurate absolute free energies for ligand-protein binding based on non-equilibrium approachesGapsys, Vytautas; Yildirim, Ahmet; Aldeghi, Matteo; Khalak, Yuriy; van der Spoel, David; de Groot, Bert L.Communications Chemistry (2021), 4 (1), 61CODEN: CCOHCT; ISSN:2399-3669. (Nature Research)Abstr.: The accurate calcn. of the binding free energy for arbitrary ligand-protein pairs is a considerable challenge in computer-aided drug discovery. Recently, it has been demonstrated that current state-of-the-art mol. dynamics (MD) based methods are capable of making highly accurate predictions. Conventional MD-based approaches rely on the first principles of statistical mechanics and assume equil. sampling of the phase space. In the current work we demonstrate that accurate abs. binding free energies (ABFE) can also be obtained via theor. rigorous non-equil. approaches. Our investigation of ligands binding to bromodomains and T4 lysozyme reveals that both equil. and non-equil. approaches converge to the same results. The non-equil. approach achieves the same level of accuracy and convergence as an equil. free energy perturbation (FEP) method enhanced by Hamiltonian replica exchange. We also compare uni- and bi-directional non-equil. approaches and demonstrate that considering the work distributions from both forward and reverse directions provides substantial accuracy gains. In summary, non-equil. ABFE calcns. are shown to yield reliable and well-converged ests. of protein-ligand binding affinity.
- 29Xu, X.; Angioletti-Uberti, S.; Lu, Y.; Dzubiella, J.; Ballauff, M. Interaction of proteins with polyelectrolytes: Comparison of theory to experiment. Langmuir 2019, 35, 5373– 5391, DOI: 10.1021/acs.langmuir.8b0180229https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhsVOltrvM&md5=843d9f0f903cbc530b2dcc43ff6118a6Interaction of Proteins with Polyelectrolytes: Comparison of Theory to ExperimentXu, Xiao; Angioletti-Uberti, Stefano; Lu, Yan; Dzubiella, Joachim; Ballauff, MatthiasLangmuir (2019), 35 (16), 5373-5391CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)A review on the authors' recent studies on the interaction of simple proteins such as human serum albumin (HSA) and lysozyme with linear polyelectrolytes, charged dendrimers, charged networks, and polyelectrolyte brushes, based on exptl. works combined with mol. dynamics (MD) simulations and mean-field theories.
- 30Paiardi, G.; Ferraz, M.; Rusnati, M.; Wade, R. C. The accomplices: Heparan sulfates and N-glycans foster SARS-CoV-2 spike: ACE2 receptor binding and virus priming. Proc. Natl. Acad. Sci. U.S.A. 2024, 121, e2404892121 DOI: 10.1073/pnas.2404892121There is no corresponding record for this reference.
- 31Bhatia, S.; Lauster, D.; Bardua, M.; Ludwig, K.; Angioletti-Uberti, S.; Popp, N.; Hoffmann, U.; Paulus, F.; Budt, M.; Stadtmüller, M. Linear polysialoside outperforms dendritic analogs for inhibition of influenza virus infection in vitro and in vivo. Biomaterials 2017, 138, 22– 34, DOI: 10.1016/j.biomaterials.2017.05.02831https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXosVWnsL8%253D&md5=409628a080116660d39f1a1a41cd6c7bLinear polysialoside outperforms dendritic analogs for inhibition of influenza virus infection in vitro and in vivoBhatia, Sumati; Lauster, Daniel; Bardua, Markus; Ludwig, Kai; Angioletti-Uberti, Stefano; Popp, Nicole; Hoffmann, Ute; Paulus, Florian; Budt, Matthias; Stadtmueller, Marlena; Wolff, Thorsten; Hamann, Alf; Boettcher, Christoph; Herrmann, Andreas; Haag, RainerBiomaterials (2017), 138 (), 22-34CODEN: BIMADU; ISSN:0142-9612. (Elsevier Ltd.)Inhibition of influenza A virus infection by multivalent sialic acid inhibitors preventing viral hemagglutinin binding to host cells of the respiratory tract is a promising strategy. However, optimal geometry and optimal ligand presentation on multivalent scaffolds for efficient inhibition both in vitro and in vivo application are still unclear. Here, by comparing linear and dendritic polyglycerol sialosides (LPGSA and dPGSA) we identified architectural requirements and optimal ligand densities for an efficient multivalent inhibitor of influenza virus A/X31/1 (H3N2). Due to its large vol., the LPGSA at optimal ligand d. sterically shielded the virus significantly better than the dendritic analog. A statistical mechanics model rationalizes the relevance of ligand d., morphol., and the size of multivalent scaffolds for the potential to inhibit virus-cell binding. Optimized LPGSA inhibited virus infection at IC50 in the low nanomolar nanoparticle concn. range and also showed potent antiviral activity against two avian influenza strains A/Mallard/439/2004 (H3N2) and A/turkey/Italy/472/1999 (H7N1) post infection. In vivo application of inhibitors clearly confirmed the higher inhibition potential of linear multivalent scaffolds to prevent infection. The optimized LPGSA did not show any acute toxicity, and was much more potent than the neuraminidase inhibitor oseltamivir carboxylate in vivo. Combined application of the LPGSA and oseltamivir carboxylate revealed a synergistic inhibitory effect and successfully prevented influenza virus infection in mice.
- 32Xu, C.; Wang, Y.; Liu, C.; Zhang, C.; Han, W.; Hong, X.; Wang, Y.; Hong, Q.; Wang, S.; Zhao, Q. Conformational dynamics of SARS-CoV-2 trimeric spike glycoprotein in complex with receptor ACE2 revealed by cryo-EM. Sci. Adv. 2021, 7, eabe5575 DOI: 10.1126/sciadv.abe5575There is no corresponding record for this reference.
- 33Horinek, D.; Serr, A.; Geisler, M.; Pirzer, T.; Slotta, U.; Lud, S. Q.; Garrido, J.; Scheibel, T.; Hugel, T.; Netz, R. R. Peptide adsorption on a hydrophobic surface results from an interplay of solvation, surface, and intrapeptide forces. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 2842– 2847, DOI: 10.1073/pnas.070787910533https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXjtVSju70%253D&md5=8509147b175b299592be205938e8d48dPeptide adsorption on a hydrophobic surface results from an interplay of solvation, surface, and intrapeptide forcesHorinek, D.; Serr, A.; Geisler, M.; Pirzer, T.; Slotta, U.; Lud, S. Q.; Garrido, J. A.; Scheibel, T.; Hugel, T.; Netz, R. R.Proceedings of the National Academy of Sciences of the United States of America (2008), 105 (8), 2842-2847CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)The hydrophobic effect, i.e., the poor solvation of nonpolar parts of mols., plays a key role in protein folding and more generally for mol. self-assembly and aggregation in aq. media. The perturbation of the water structure accounts for many aspects of protein hydrophobicity. However, to what extent the dispersion interaction between mol. entities themselves contributes has remained unclear. This is so because in peptide folding interactions and structural changes occur on all length scales and make disentangling various contributions impossible. The authors address this issue both exptl. and theor. by looking at the force necessary to peel a mildly hydrophobic single peptide mol. from a flat hydrophobic diamond surface in the presence of water. This setup avoids problems caused by bubble adsorption, cavitation, and slow equilibration that complicate the much-studied geometry with two macroscopic surfaces. Using at.-force spectroscopy, the authors det. the mean desorption force of a single spider-silk peptide chain as F = 58 ± 8 pN, which corresponds to a desorption free energy of ≈5 k8T per amino acid. The authors' all-atomistic mol. dynamics simulation including explicit water correspondingly yields the desorption force F = 54 ± 15 pN. This observation demonstrates that std. nonpolarizable force fields used in classical simulations are capable of resolving the fine details of the hydrophobic attraction of peptides. The anal. of the involved energetics shows that water-structure effects and dispersive interactions give contributions of comparable magnitude that largely cancel out. It follows that the correct modeling of peptide hydrophobicity must take the intimate coupling of solvation and dispersive effects into account.
- 34Schwierz, N.; Horinek, D.; Liese, S.; Pirzer, T.; Balzer, B. N.; Hugel, T.; Netz, R. R. On the relationship between peptide adsorption resistance and surface contact angle: a combined experimental and simulation single-molecule study. J. Am. Chem. Soc. 2012, 134, 19628– 19638, DOI: 10.1021/ja304462u34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhsFygtb%252FJ&md5=ebb99ca67cd645f7d3553cf517a54ed3On the Relationship between Peptide Adsorption Resistance and Surface Contact Angle: A Combined Experimental and Simulation Single-Molecule StudySchwierz, Nadine; Horinek, Dominik; Liese, Susanne; Pirzer, Tobias; Balzer, Bizan N.; Hugel, Thorsten; Netz, Roland R.Journal of the American Chemical Society (2012), 134 (48), 19628-19638CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The force-induced desorption of single peptide chains from mixed OH/CH3-terminated self-assembled monolayers was studied in closely matched mol. dynamics simulations and at. force microscopy expts. with the goal to gain microscopic understanding of the transition between peptide adsorption and adsorption resistance as the surface contact angle is varied. In both simulations and expts., the surfaces become adsorption resistant against hydrophilic as well as hydrophobic peptides when their contact angle decreases below θ ≈ 50°-60°, thus confirming the so-called Berg limit established in the context of protein and cell adsorption. Entropy/enthalpy decompn. of the simulation results reveals that the key discriminator between the adsorption of different residues on a hydrophobic monolayer is of entropic nature and thus probably is linked to the hydrophobic effect. By pushing a polyalanine peptide onto a polar surface, simulations reveal that the peptide adsorption resistance is caused by the strongly bound water hydration layer and characterized by the simultaneous gain of both total entropy in the system and total no. of hydrogen bonds between water, peptide, and surface. This mechanistic insight into peptide adsorption resistance might help to refine design principles for anti-fouling surfaces.
- 35Page, T. M.; Nie, C.; Neander, L.; Povolotsky, T. L.; Sahoo, A. K.; Nickl, P.; Adler, J. M.; Bawadkji, O.; Radnik, J.; Achazi, K. Functionalized Fullerene for Inhibition of SARS-CoV-2 Variants. Small 2023, 19, 2206154 DOI: 10.1002/smll.20220615435https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3sXhsFylur0%253D&md5=ca14d876275ed6fe55ed8d23d417e1e6Functionalized Fullerene for Inhibition of SARS-CoV-2 VariantsPage, Taylor M.; Nie, Chuanxiong; Neander, Lenard; Povolotsky, Tatyana L.; Sahoo, Anil Kumar; Nickl, Philip; Adler, Julia M.; Bawadkji, Obida; Radnik, Jorg; Achazi, Katharina; Ludwig, Kai; Lauster, Daniel; Netz, Roland R.; Trimpert, Jakob; Kaufer, Benedikt; Haag, Rainer; Donskyi, Ievgen S.Small (2023), 19 (15), 2206154CODEN: SMALBC; ISSN:1613-6810. (Wiley-VCH Verlag GmbH & Co. KGaA)As virus outbreaks continue to pose a challenge, a nonspecific viral inhibitor can provide significant benefits, esp. against respiratory viruses. Polyglycerol sulfates recently emerge as promising agents that mediate interactions between cells and viruses through electrostatics, leading to virus inhibition. Similarly, hydrophobic C60 fullerene can prevent virus infection via interactions with hydrophobic cavities of surface proteins. Here, two strategies are combined to inhibit infection of SARS-CoV-2 variants in vitro. Effective inhibitory concns. in the millimolar range highlight the significance of bare fullerene's hydrophobic moiety and electrostatic interactions of polysulfates with surface proteins of SARS-CoV-2. Furthermore, microscale thermophoresis measurements support that fullerene linear polyglycerol sulfates interact with the SARS-CoV-2 virus via its spike protein, and highlight importance of electrostatic interactions within it. All-atom mol. dynamics simulations reveal that the fullerene binding site is situated close to the receptor binding domain, within 4 nm of polyglycerol sulfate binding sites, feasibly allowing both portions of the material to interact simultaneously.
- 36Woo, H.; Park, S.-J.; Choi, Y. K.; Park, T.; Tanveer, M.; Cao, Y.; Kern, N. R.; Lee, J.; Yeom, M. S.; Croll, T. I. Developing a fully glycosylated full-length SARS-CoV-2 spike protein model in a viral membrane. J. Phys. Chem. B 2020, 124, 7128– 7137, DOI: 10.1021/acs.jpcb.0c0455336https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXht1WmsrjK&md5=a74433db7db78137b36d5fd714d49dd1Developing a fully glycosylated full-length SARS-CoV-2 spike protein model in a viral membraneWoo, Hyeonuk; Park, Sang-Jun; Choi, Yeol Kyo; Park, Taeyong; Tanveer, Maham; Cao, Yiwei; Kern, Nathan R.; Lee, Jumin; Yeom, Min Sun; Croll, Tristan I.; Seok, Chaok; Im, WonpilJournal of Physical Chemistry B (2020), 124 (33), 7128-7137CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)This tech. study describes all-atom modeling and simulation of a fully glycosylated full-length SARS-CoV-2 spike (S) protein in a viral membrane. First, starting from PDB: 6VSB and 6VXX, full-length S protein structures were modeled using template-based modeling, de-novo protein structure prediction, and loop modeling techniques in GALAXY modeling suite. Then, using the recently detd. most occupied glycoforms, 22 N-glycans and 1 O-glycan of each monomer were modeled using Glycan Reader & Modeler in CHARMM-GUI. These fully glycosylated full-length S protein model structures were assessed and further refined against the low-resoln. data in their resp. exptl. maps using ISOLDE. The authors then used CHARMM-GUI Membrane Builder to place the S proteins in a viral membrane and performed all-atom mol. dynamics simulations. All structures are available in CHARMM-GUI COVID-19 Archive so that researchers can use these models to carry out innovative and novel modeling and simulation research for the prevention and treatment of COVID-19.
- 37Casalino, L.; Gaieb, Z.; Goldsmith, J. A.; Hjorth, C. K.; Dommer, A. C.; Harbison, A. M.; Fogarty, C. A.; Barros, E. P.; Taylor, B. C.; McLellan, J. S. Beyond shielding: the roles of glycans in the SARS-CoV-2 spike protein. ACS Cent. Sci. 2020, 6, 1722– 1734, DOI: 10.1021/acscentsci.0c0105637https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhvVOlsb3N&md5=52d499afcd7e3caa7d9e6017ffa86e45Beyond Shielding: The Roles of Glycans in the SARS-CoV-2 Spike ProteinCasalino, Lorenzo; Gaieb, Zied; Goldsmith, Jory A.; Hjorth, Christy K.; Dommer, Abigail C.; Harbison, Aoife M.; Fogarty, Carl A.; Barros, Emilia P.; Taylor, Bryn C.; McLellan, Jason S.; Fadda, Elisa; Amaro, Rommie E.ACS Central Science (2020), 6 (10), 1722-1734CODEN: ACSCII; ISSN:2374-7951. (American Chemical Society)The ongoing COVID-19 pandemic caused by severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) has resulted in more than 28,000,000 infections and 900,000 deaths worldwide to date. Antibody development efforts mainly revolve around the extensively glycosylated SARS-CoV-2 spike (S) protein, which mediates host cell entry by binding to the angiotensin-converting enzyme 2 (ACE2). Similar to many other viral fusion proteins, the SARS-CoV-2 spike utilizes a glycan shield to thwart the host immune response. Here, we built a full-length model of the glycosylated SARS-CoV-2 S protein, both in the open and closed states, augmenting the available structural and biol. data. Multiple microsecond-long, all-atom mol. dynamics simulations were used to provide an atomistic perspective on the roles of glycans and on the protein structure and dynamics. We reveal an essential structural role of N-glycans at sites N165 and N234 in modulating the conformational dynamics of the spike's receptor binding domain (RBD), which is responsible for ACE2 recognition. This finding is corroborated by biolayer interferometry expts., which show that deletion of these glycans through N165A and N234A mutations significantly reduces binding to ACE2 as a result of the RBD conformational shift toward the "down" state. Addnl., end-to-end accessibility analyses outline a complete overview of the vulnerabilities of the glycan shield of the SARS-CoV-2 S protein, which may be exploited in the therapeutic efforts targeting this mol. machine. Overall, this work presents hitherto unseen functional and structural insights into the SARS-CoV-2 S protein and its glycan coat, providing a strategy to control the conformational plasticity of the RBD that could be harnessed for vaccine development. The glycan shield is a sugary barrier that helps the viral SARS-CoV-2 spikes to evade the immune system. Beyond shielding, two of the spike's glycans are discovered to prime the virus for infection.
- 38Erbaş, A.; Netz, R. R. Confinement-dependent friction in peptide bundles. Biophys. J. 2013, 104, 1285– 1295, DOI: 10.1016/j.bpj.2013.02.00838https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXktlKrs78%253D&md5=ccca5f67247a960798d12d28893b7797Confinement-Dependent Friction in Peptide BundlesErbas, Aykut; Netz, Roland R.Biophysical Journal (2013), 104 (6), 1285-1295CODEN: BIOJAU; ISSN:0006-3495. (Cell Press)Friction within globular proteins or between adhering macromols. crucially dets. the kinetics of protein folding, the formation, and the relaxation of self-assembled mol. systems. One fundamental question is how these friction effects depend on the local environment and in particular on the presence of water. In this model study, we use fully atomistic MD simulations with explicit water to obtain friction forces as a single polyglycine peptide chain is pulled out of a bundle of k adhering parallel polyglycine peptide chains. The whole system is periodically replicated along the peptide axes, so a stationary state at prescribed mean sliding velocity V is achieved. The aggregation no. is varied between k = 2 (two peptide chains adhering to each other with plenty of water present at the adhesion sites) and k = 7 (one peptide chain pulled out from a close-packed cylindrical array of six neighboring peptide chains with no water inside the bundle). The friction coeff. per hydrogen bond, extrapolated to the viscous limit of vanishing pulling velocity V → 0, exhibits an increase by five orders of magnitude when going from k = 2 to k = 7. This dramatic confinement-induced friction enhancement we argue to be due to a combination of water depletion and increased hydrogen-bond cooperativity.
- 39Erbaş, A.; Horinek, D.; Netz, R. R. Viscous friction of hydrogen-bonded matter. J. Am. Chem. Soc. 2012, 134, 623– 630, DOI: 10.1021/ja209454a39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhsV2jtb3K&md5=7fa0d24b3bd97c4821f9ba0f4c05aa3aViscous Friction of Hydrogen-Bonded MatterErbas, Aykut; Horinek, Dominik; Netz, Roland R.Journal of the American Chemical Society (2012), 134 (1), 623-630CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Amontons' law successfully describes friction between macroscopic solid bodies for a wide range of velocities and normal forces. For the diffusion and forced sliding of adhering or entangled macromols., proteins, and biol. complexes, temp. effects are invariably important, and a similarly successful friction law at biol. length and velocity scales is missing. Hydrogen bonds (HBs) are key to the specific binding of biomatter. Here we show that friction between hydrogen-bonded matter in the biol. relevant low-velocity viscous regime obeys a simple law: the friction force is proportional to the no. of HBs, the sliding velocity, and a friction coeff. γHB. This law is deduced from atomistic mol. dynamics simulations for short peptide chains that are laterally pulled over planar hydroxylated substrates in the presence of water and holds for widely different peptides, surface polarities, and applied normal forces. The value of γHB is extrapolated from simulations at sliding velocities in the range from V = 10-2 to 100 m/s by mapping on a simple stochastic model and turns out to be of the order of γHB ≃ 10-8 kg/s. The friction of a single HB thus amts. to the Stokes friction of a sphere with an equiv. radius of roughly 1 μm moving in water. Cooperativity is pronounced; roughly three HBs act collectively.
- 40Patil, S. P.; Xiao, S.; Gkagkas, K.; Markert, B.; Gräter, F. Viscous friction between crystalline and amorphous phase of dragline silk. PLoS One 2014, 9, e104832 DOI: 10.1371/journal.pone.0104832There is no corresponding record for this reference.
- 41Kumar, S.; Rosenberg, J. M.; Bouzida, D.; Swendsen, R. H.; Kollman, P. A. The weighted histogram analysis method for free-energy calculations on biomolecules. I. The method. J. Comput. Chem. 1992, 13, 1011– 1021, DOI: 10.1002/jcc.54013081241https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK38XmtVynsrs%253D&md5=5b2ad7410198f03025708a37c0fbe89dThe weighted histogram analysis method for free-energy calculations on biomolecules. I. The methodKumar, Shankar; Bouzida, Djamal; Swendsen, Robert H.; Kollman, Peter A.; Rosenberg, John M.Journal of Computational Chemistry (1992), 13 (8), 1011-21CODEN: JCCHDD; ISSN:0192-8651.The Weighted Histogram Anal. Method (WHAM), an extension of Ferrenberg and Swendsen's Multiple Histogram Technique, has been applied for the first time on complex biomol. Hamiltonians. The method is presented here as an extension of the Umbrella Sampling method for free-energy and Potential of Mean Force calcns. This algorithm possesses the following advantages over methods that are currently employed: (1) it provides a built-in est. of sampling errors thereby yielding objective ests. of the optimal location and length of addnl. simulations needed to achieve a desired level of precision; (2) it yields the "best" value of free energies by taking into account all the simulations so as to minimize the statistical errors; (3) in addn. to optimizing the links between simulations, it also allows multiple overlaps of probability distributions for obtaining better ests. of the free-energy differences. By recasting the Ferrenberg-Swendsen Multiple Histogram equations in a form suitable for mol. mechanics type Hamiltonians, we have demonstrated the feasibility and robustness of this method by applying it to a test problem of the generation of the Potential of Mean Force profile of the pseudorotation phase angle of the sugar ring in deoxyadenosine.
- 42Torrie, G. M.; Valleau, J. P. Nonphysical sampling distributions in Monte Carlo free-energy estimation: Umbrella sampling. J. Comput. Phys. 1977, 23, 187– 199, DOI: 10.1016/0021-9991(77)90121-8There is no corresponding record for this reference.
- 43Phillips, R.; Kondev, J.; Theriot, J.; Garcia, H. Physical Biology of the Cell, 2nd ed.; Garland Science: New York, 2012; p 1088.There is no corresponding record for this reference.
- 44Hanke, F.; Serr, A.; Kreuzer, H. J.; Netz, R. R. Stretching single polypeptides: The effect of rotational constraints in the backbone. Europhys. Lett. 2010, 92, 53001 DOI: 10.1209/0295-5075/92/53001There is no corresponding record for this reference.
- 45Kitov, P. I.; Bundle, D. R. On the nature of the multivalency effect: a thermodynamic model. J. Am. Chem. Soc. 2003, 125, 16271– 16284, DOI: 10.1021/ja038223n45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXpsFOkuro%253D&md5=9ba6a3a7af88f5be5c790b32eb73aee1On the Nature of the Multivalency Effect: A Thermodynamic ModelKitov, Pavel I.; Bundle, David R.Journal of the American Chemical Society (2003), 125 (52), 16271-16284CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)A quant. model is proposed for the anal. of the thermodn. parameters of multivalent interactions in dil. solns. or with immobilized multimeric receptor. The model takes into account all bound species and describes multivalent binding via two microscopic binding energies corresponding to inter- and intramol. interactions (ΔG°inter and ΔG°intra), the relative contributions of which depend on the distribution of complexes with different nos. of occupied binding sites. The third component of the overall free energy, which we call the "avidity entropy" term, is a function of the degeneracy of bound states, Ωi, which is calcd. on the basis of the topol. of interaction and the distribution of all bound species. This term grows rapidly with the no. of receptor sites and ligand multivalency, it always favors binding, and explains why multivalency can overcome the loss of conformational entropy when ligands displayed at the ends of long tethers are bound. The microscopic parameters ΔG°inter and ΔG°intra may be detd. from the obsd. binding energies for a set of oligovalent ligands by nonlinear fitting with the theor. model. Here binding data obtained from two series of oligovalent carbohydrate inhibitors for Shiga-like toxins were used to verify the theory. The decavalent and octavalent inhibitors exhibit subnanomolar activity and are the most active sol. inhibitors yet seen that block Shiga-like toxin binding to its native receptor. The theory developed here in conjunction with our protocol for the optimization of tether length provides a predictive approach to design and maximize the avidity of multivalent ligands.
- 46Zumbro, E.; Witten, J.; Alexander-Katz, A. Computational insights into avidity of polymeric multivalent binders. Biophys. J. 2019, 117, 892– 902, DOI: 10.1016/j.bpj.2019.07.02646https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhsFemtb3L&md5=af2eb7445718e750eebf33397009c4aeComputational Insights into Avidity of Polymeric Multivalent BindersZumbro, Emiko; Witten, Jacob; Alexander-Katz, AlfredoBiophysical Journal (2019), 117 (5), 892-902CODEN: BIOJAU; ISSN:0006-3495. (Cell Press)Multivalent binding interactions are commonly found throughout biol. to enhance weak monovalent binding such as between glycoligands and protein receptors. Designing multivalent polymers to bind to viruses and toxic proteins is a promising avenue for inhibiting their attachment and subsequent infection of cells. Several studies have focused on oligomeric multivalent inhibitors and how changing parameters such as ligand shape, size, linker length, and flexibility affect binding. However, exptl. studies of how larger structural parameters of multivalent polymers, such as d.p., affect binding avidity to targets have mixed results, with some finding an improvement with longer polymers and some finding no effect. Here, we use Brownian dynamics simulations to provide a theor. understanding of how the d.p. affects the binding avidity of multivalent polymers. We show that longer polymers increase binding avidity to multivalent targets but reach a limit in binding avidity at high ds.p. We also show that when interacting with multiple targets simultaneously, longer polymers are able to use intertarget interactions to promote clustering and improve binding efficiency. We expect our results to narrow the design space for optimizing the structure and effectiveness of multivalent inhibitors as well as be useful to understand biol. design strategies for multivalent binding.
- 47Qiao, B.; Jiménez-Ángeles, F.; Nguyen, T. D.; Olvera de la Cruz, M. Water follows polar and nonpolar protein surface domains. Proc. Natl. Acad. Sci. U.S.A. 2019, 116, 19274– 19281, DOI: 10.1073/pnas.191022511647https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhvVGgu7jP&md5=cec3762bbfde8ac397491e088c567a9eWater follows polar and nonpolar protein surface domainsQiao, Baofu; Jimenez-Angeles, Felipe; Nguyen, Trung Dac; de la Cruz, Monica OlveraProceedings of the National Academy of Sciences of the United States of America (2019), 116 (39), 19274-19281CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)The conformation of water around proteins is of paramount importance, as it dets. protein interactions. Although the av. water properties around the surface of proteins have been provided exptl. and computationally, protein surfaces are highly heterogeneous. Therefore, it is crucial to det. the correlations of water to the local distributions of polar and nonpolar protein surface domains to understand functions such as aggregation, mutations, and delivery. By using atomistic simulations, we investigate the orientation and dynamics of water mols. next to 4 types of protein surface domains: neg. charged, pos. charged, and charge-neutral polar and nonpolar amino acids. The neg. charged amino acids orient around 98% of the neighboring water dipoles toward the protein surface, and such correlation persists up to around 16 Å from the protein surface. The pos. charged amino acids orient around 94% of the nearest water dipoles against the protein surface, and the correlation persists up to around 12 Å. The charge-neutral polar and nonpolar amino acids are also orienting the water neighbors in a quant. weaker manner. A similar trend was obsd. in the residence time of the nearest water neighbors. These findings hold true for 3 tech. important enzymes (PETase, cytochrome P 450, and organophosphorus hydrolase). Our results demonstrate that the water-amino acid degree of correlation follows the same trend as the amino acid contribution in proteins soly., namely, the neg. charged amino acids are the most beneficial for protein soly., then the pos. charged amino acids, and finally the charge-neutral amino acids.
- 48Manning, G. S. Limiting laws and counterion condensation in polyelectrolyte solutions I. Colligative properties. J. Chem. Phys. 1969, 51, 924– 933, DOI: 10.1063/1.167215748https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaF1MXkslCrsLk%253D&md5=9e5de22424890b7e855a638456724ddfLimiting laws and counterion condensation in polyelectroyte solutions. I. Colligative propertiesManning, Gerald S.Journal of Chemical Physics (1969), 51 (3), 924-33CODEN: JCPSA6; ISSN:0021-9606.Formulas are derived for the osmotic coeff., the Donnan salt-exclusion factor, and the mobile-ion activity coeffs. in a polyelectrolyte soln. with or without added sample salt. The formulas, which contain no adjustable parameters, are based on the (theoretical) observation by several workers that counterions will "condense" on the polyion until the charge d. on the polyion is reduced below a certain crit. value. The uncondensed mobile ions are treated in the Debye-Hueckel approxn. In a restricted sense, the formulas are "limiting laws," and this aspect is discussed at length. Detailed comparison with exptl. data in the literature is given; agreement of the theory with expt. is usually quant.
- 49Fenley, M. O.; Manning, G. S.; Olson, W. K. Approach to the limit of counterion condensation. Biopolymers 1990, 30, 1191– 1203, DOI: 10.1002/bip.36030130549https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3MXhvVSmsLg%253D&md5=0370802614aafca9380719feb02c39a1Approach to the limit of counterion condensationFenley, Marcia O.; Manning, Gerald S.; Olson, Wilma K.Biopolymers (1990), 30 (13-14), 1191-203CODEN: BIPMAA; ISSN:0006-3525.According to counterion condensation theory, one of the contributions to the polyelectrolyte free energy is a pairwise sum of Debye-Hueckel potentials between polymer charges that are reduced by condensed counterions. When the polyion model is taken as an infinitely long and uniformly spaced line of charges, a simple closed expression for the summation, combined with entropy-derived mixing contributions, leads to the central result of the theory, a condensed fraction of counterions dependent only on the linear charge d. of the polyion and the valence of the counterion, stable against increases of salt up to concns. in excess of 0.1M. Here the sum is numerically evaluated for B-DNA models other than the infinite line of B-DNA charges. For a finite-length line there are end effects at low salt. The condensation limit is reached as a flat plateau by increasing the salt concn. At a fixed salt concn. the condensation limit is reached by increasing the length of the line. At moderate salt even very short B-DNA line-model oligomers have condensed fractions not far from the infinite polymer limit. For a long double-helical array with charge coordinates at the phosphates of B-DNA, the limiting condensed fraction appears to be approached at low salt. In contrast to the results for the line of charges, however, the computed condensed fraction varies strongly with salt in the range of exptl. typical concns. Salt invariance is restored, in agreement with both the line model and exptl. data, when dielec. satn. is considered by means of a distance-dependent dielec. function. For sufficiently long B-DNA line and helical models, at typical salt concns., the counterion binding fraction approaches the polymer limit as a linear function of 1/P, where P is the no. of phosphate groups of B-DNA.
- 50Walkowiak, J. J.; Ballauff, M. Interaction of polyelectrolytes with proteins: quantifying the role of water. Adv. Sci. 2021, 8, 2100661 DOI: 10.1002/advs.20210066150https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXisV2qu7bM&md5=78e86f5b8efd62947c1f047b2890347fInteraction of Polyelectrolytes with Proteins: Quantifying the Role of WaterWalkowiak, Jacek J.; Ballauff, MatthiasAdvanced Science (Weinheim, Germany) (2021), 8 (12), 2100661CODEN: ASDCCF; ISSN:2198-3844. (Wiley-VCH Verlag GmbH & Co. KGaA)A theor. model is presented for the free energy ΔGb of complex formation between a highly charged polyelectrolyte and a protein. The model introduced here comprises both the effect of released counterions and the uptake or release of water mols. during complex formation. The resulting expression for ΔGb is hence capable of describing the dependence of ΔGb on temp. as well as on the concn. of salt in the system: An increase of the salt concn. in the soln. increases the activity of the ions and counterion release becomes less effective for binding. On the other hand, an increased salt concn. leads to the decrease of the activity of water in bulk. Hence, release of water mols. during complex formation will be more advantageous and lead to an increase of the magnitude of ΔGb and the binding const. It is furthermore demonstrated that the release or uptake of water mols. is the origin of the marked enthalpy-entropy cancellation obsd. during complex formation of polyelectrolytes with proteins. The comparison with exptl. data on complex formation between a synthetic (sulfated dendritic polyglycerol) and natural polyelectrolytes (DNA; heparin) with proteins shows full agreement with theory.
- 51Irudayam, S. J.; Henchman, R. H. Entropic cost of protein- ligand binding and its dependence on the entropy in solution. J. Phys. Chem. B 2009, 113, 5871– 5884, DOI: 10.1021/jp809968p51https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXktVOltLY%253D&md5=997977e2ace284e3af8e6f46ae347dd0Entropic Cost of Protein-Ligand Binding and Its Dependence on the Entropy in SolutionIrudayam, Sheeba Jem; Henchman, Richard H.Journal of Physical Chemistry B (2009), 113 (17), 5871-5884CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)Two theor. formulations are proposed and compared for the loss of translational and rotational entropy upon protein-ligand binding in water. The two theories share the same approach to evaluate the translational and rotational entropy of the ligand when bound. The potential of the bound ligand is modeled by six harmonic oscillators that are parametrized from the force and torque magnitudes measured in a mol. dynamics simulation, yielding vibrational and librational entropies. In the aq. phase, the theories differ because there is no unique way to assign the total entropy to mols. in soln. In one approach, the ligand is allowed unrestricted access to the full soln. vol. at the std. concn. and is assigned the same translational and rotational entropy as if it were an ideal gas. We term this a "mol.-frame" (MF) theory because it considers configurational space in the ref. frame of the mol. of interest. The entropy of the solvent is penalized because it is excluded from the mol.'s vol. In the second theory, all mols. including the solvent are confined by their neighbors in mean-field configurational vols. This we term a "system-frame" (SF) theory because the configurational space available to all mols. is considered in the ref. frame of the whole system. Mols. have vibrational and librational entropy in the same way as they do when bound. In addn., the discrete size of the solvent mols. quantizes the configurational space into an effective no. of min. according to the solute mol.'s std. concn. and the mean vol. of a solvent mol. This leads to the cratic entropy expressed in terms of the solute mol.'s mole fraction. The equivalent no. of min. in rotational space depends on both the solute mol.'s vol. and the solvent mol.'s vol. This leads to an equation for the orientational entropy based on the proposed concept of "angle fraction". The MF and SF theories are applied to calc. the translational and rotational entropy losses involved in the formation of six different protein-ligand complexes, in two of which the ligand is water. The MF entropy losses range from -80 to -142 J K-1 mol-1 for ligands at the 1 M std.-state concn. and from -52 to -63 J K-1 mol-1 for water at the 55.6 M std.-state concn. They depend logarithmically on both the no. and strength of interactions between the ligand and protein through the forces and torques. This is obsd. to lead to moderate dependencies on the ligands' moments of inertia and masses. The SF entropy losses are smaller and range from -50 to -75 J K-1 mol-1 for ligands at the 1 M std.-state concn. and from 0 to -12 J K-1 mol-1 for water. They depend logarithmically on the ligand solvent's mol. vol. and weakly on the relative strengths of the ligand's interactions with the protein and water. The cratic entropy loss in water at the std. concn. is const. and is also demonstrated to be implicit in MF theories. Entropy losses from the two approaches are also compared with those from other computational approaches and with expt. The use of the force and torque magnitudes leads to smaller bound vols. than are obtained from ligand-displacement approaches. The general agreement of the SF entropy losses with those from expt. suggests that the SF theory is more consistent with the assumptions made in exptl. measurements than the MF solvation theories, which would require a compensating entropy gain in the solvent in order to agree.
- 52Xu, X.; Ran, Q.; Dey, P.; Nikam, R.; Haag, R.; Ballauff, M.; Dzubiella, J. Counterion-release entropy governs the inhibition of serum proteins by polyelectrolyte drugs. Biomacromolecules 2018, 19, 409– 416, DOI: 10.1021/acs.biomac.7b0149952https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXitVWgurvN&md5=82e6f191e2e59bdd495800992841eca1Counterion-Release Entropy Governs the Inhibition of Serum Proteins by Polyelectrolyte DrugsXu, Xiao; Ran, Qidi; Dey, Pradip; Nikam, Rohit; Haag, Rainer; Ballauff, Matthias; Dzubiella, JoachimBiomacromolecules (2018), 19 (2), 409-416CODEN: BOMAF6; ISSN:1525-7797. (American Chemical Society)Dendritic polyelectrolytes constitute high potential drugs and carrier systems for biomedical purposes. Still, their biomol. interaction modes, in particular those detg. the binding affinity to proteins, have not been rationalized. We study the interaction of the drug candidate dendritic polyglycerol sulfate (dPGS) with serum proteins using isothermal titrn. calorimetry (ITC) interpreted and complemented with mol. computer simulations. Lysozyme is first studied as a well-defined model protein to verify theor. concepts, which are then applied to the important cell adhesion protein family of selectins. We demonstrate that the driving force of the strong complexation, leading to a distinct protein corona, originates mainly from the release of only a few condensed counterions from the dPGS upon binding. The binding const. shows a surprisingly weak dependence on dPGS size (and bare charge) which can be understood by colloidal charge-renormalization effects and by the fact that the magnitude of the dominating counterion-release mechanism almost exclusively depends on the interfacial charge structure of the protein-specific binding patch. Our findings explain the high selectivity of P- and L-selectins over E-selectin for dPGS to act as a highly anti-inflammatory drug. The entire anal. demonstrates that the interaction of proteins with charged polymeric drugs can be predicted by simulations with unprecedented accuracy. Thus, our results open new perspectives for the rational design of charged polymeric drugs and carrier systems.
- 53Caro, J. A.; Harpole, K. W.; Kasinath, V.; Lim, J.; Granja, J.; Valentine, K. G.; Sharp, K. A.; Wand, A. J. Entropy in molecular recognition by proteins. Proc. Natl. Acad. Sci. U.S.A. 2017, 114, 6563– 6568, DOI: 10.1073/pnas.162115411453https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXpt1Onsbo%253D&md5=7fc266b399633ec8d63e12bce098e228Entropy in molecular recognition by proteinsCaro, Jose A.; Harpole, Kyle W.; Kasinath, Vignesh; Lim, Jackwee; Granja, Jeffrey; Valentine, Kathleen G.; Sharp, Kim A.; Wand, A. JoshuaProceedings of the National Academy of Sciences of the United States of America (2017), 114 (25), 6563-6568CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)Mol. recognition by proteins is fundamental to mol. biol. Dissection of the thermodn. energy terms governing protein-ligand interactions has proven difficult, with detn. of entropic contributions being particularly elusive. NMR relaxation measurements have suggested that changes in protein conformational entropy can be quant. obtained through a dynamical proxy, but the generality of this relationship has not been shown. Twenty-eight protein-ligand complexes are used to show a quant. relationship between measures of fast side-chain motion and the underlying conformational entropy. We find that the contribution of conformational entropy can range from favorable to unfavorable, which demonstrates the potential of this thermodn. variable to modulate protein-ligand interactions. For about one-quarter of these complexes, the absence of conformational entropy would render the resulting affinity biol. meaningless. The dynamical proxy for conformational entropy or "entropy meter" also allows for refinement of the contributions of solvent entropy and the loss in rotational-translational entropy accompanying formation of high-affinity complexes. Furthermore, structure-based application of the approach can also provide insight into long-lived specific water-protein interactions that escape the generic treatments of solvent entropy based simply on changes in accessible surface area. These results provide a comprehensive and unified view of the general role of entropy in high-affinity mol. recognition by proteins.
- 54Sahoo, A. K.; Schreiber, F.; Netz, R. R.; Maiti, P. K. Role of entropy in determining the phase behavior of protein solutions induced by multivalent ions. Soft Matter 2022, 18, 592– 601, DOI: 10.1039/D1SM00730KThere is no corresponding record for this reference.
- 55Rapaport, D. Hydrogen bonds in water: Network organization and lifetimes. Mol. Phys. 1983, 50, 1151– 1162, DOI: 10.1080/0026897830010293155https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2cXhvVSntrw%253D&md5=9a83628aacab20bcca277cd3ae17c203Hydrogen bonds in water. Network organization and lifetimesRapaport, D. C.Molecular Physics (1983), 50 (5), 1151-62CODEN: MOPHAM; ISSN:0026-8976.Equil. and dynamical properties of H bonds in liq. H2O are analyzed using the results of mol. dynamics simulations of the MCY-CI model. Properties of the H bond clusters as functions of temp. are described. The connectivity of the clusters is analyzed in terms of the bridgeless polygons which are formed by the bonds. The problem of obtaining a meaningful definition of bond lifetime is discussed, and the results of lifetime measurements based on alternative definitions are shown.
- 56Malicka, W.; Haag, R.; Ballauff, M. Interaction of heparin with proteins: hydration effects. J. Phys. Chem. B 2022, 126, 6250– 6260, DOI: 10.1021/acs.jpcb.2c04928There is no corresponding record for this reference.
- 57Geisler, M.; Xiao, S.; Puchner, E. M.; Gräter, F.; Hugel, T. Controlling the structure of proteins at surfaces. J. Am. Chem. Soc. 2010, 132, 17277– 17281, DOI: 10.1021/ja107212z57https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhsVeksL%252FI&md5=c7af30feab283481a7e018b07bc0892eControlling the Structure of Proteins at SurfacesGeisler, Michael; Xiao, Sen-Bo; Puchner, Elias M.; Graeter, Frauke; Hugel, ThorstenJournal of the American Chemical Society (2010), 132 (48), 17277-17281CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)With the help of single mol. force spectroscopy and mol. dynamics simulations, we det. the surface-induced structure of a single engineered spider silk protein. An amyloid like structure is induced in the vicinity of a surface with high surface energy and can be prohibited in the presence of a hydrophobic surface. The derived mol. energy landscapes highlight the role of single silk protein structure for the macroscopic toughness of spider silk.
- 58Liese, S.; Netz, R. R. Influence of length and flexibility of spacers on the binding affinity of divalent ligands. Beilstein J. Org. Chem. 2015, 11, 804– 816, DOI: 10.3762/bjoc.11.9058https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXpvFGntro%253D&md5=43e11f85192cf5716dafcc802eec77d3Influence of length and flexibility of spacers on the binding affinity of divalent ligandsLiese, Susanne; Netz, Roland R.Beilstein Journal of Organic Chemistry (2015), 11 (), 804-816CODEN: BJOCBH; ISSN:1860-5397. (Beilstein-Institut zur Foerderung der Chemischen Wissenschaften)We present a quant. model for the binding of divalent ligand-receptor systems. We study the influence of length and flexibility of the spacers on the overall binding affinity and derive general rules for the optimal ligand design. To this end, we first compare different polymeric models and det. the probability to simultaneously bind to two neighboring receptor binding pockets. In a second step the binding affinity of divalent ligands in terms of the IC50 value is derived. We find that a divalent ligand has the potential to bind more efficiently than its monovalent counterpart only, if the monovalent dissocn. const. is lower than a crit. value. This crit. monovalent dissocn. const. depends on the ligand-spacer length and flexibility as well as on the size of the receptor. Regarding the optimal ligand-spacer length and flexibility, we find that the av. spacer length should be equal or slightly smaller than the distance between the receptor binding pockets and that the end-to-end spacer length fluctuations should be in the same range as the size of a receptor binding pocket.
- 59Jo, S.; Kim, T.; Iyer, V. G.; Im, W. CHARMM-GUI: a web-based graphical user interface for CHARMM. J. Comput. Chem. 2008, 29, 1859– 1865, DOI: 10.1002/jcc.2094559https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXosVKksbc%253D&md5=112a3dd61d792b040f9f716b32220d7eCHARMM-GUI: a web-based graphical user interface for CHARMMJo, Sunhwan; Kim, Taehoon; Iyer, Vidyashankara G.; Im, WonpilJournal of Computational Chemistry (2008), 29 (11), 1859-1865CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)CHARMM is an academic research program used widely for macromol. mechanics and dynamics with versatile anal. and manipulation tools of at. coordinates and dynamics trajectories. CHARMM-GUI, http://www.charmm-gui.org, has been developed to provide a web-based graphical user interface to generate various input files and mol. systems to facilitate and standardize the usage of common and advanced simulation techniques in CHARMM. The web environment provides an ideal platform to build and validate a mol. model system in an interactive fashion such that, if a problem is found through visual inspection, one can go back to the previous setup and regenerate the whole system again. In this article, we describe the currently available functional modules of CHARMM-GUI Input Generator that form a basis for the advanced simulation techniques. Future directions of the CHARMM-GUI development project are also discussed briefly together with other features in the CHARMM-GUI website, such as Archive and Movie Gallery.
- 60Park, S.-J.; Lee, J.; Qi, Y.; Kern, N. R.; Lee, H. S.; Jo, S.; Joung, I.; Joo, K.; Lee, J.; Im, W. CHARMM-GUI Glycan Modeler for modeling and simulation of carbohydrates and glycoconjugates. Glycobiology 2019, 29, 320– 331, DOI: 10.1093/glycob/cwz00360https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXitFajsLzJ&md5=0bd9d6b408efabd1d1a1440c9a639100CHARMM-GUI Glycan Modeler for modeling and simulation of carbohydrates and glycoconjugatesPark, Sang-Jun; Lee, Jumin; Qi, Yifei; Kern, Nathan R.; Lee, Hui Sun; Jo, Sunhwan; Joung, Insuk; Joo, Keehyung; Lee, Jooyoung; Im, WonpilGlycobiology (2019), 29 (4), 320-331CODEN: GLYCE3; ISSN:1460-2423. (Oxford University Press)Characterizing glycans and glycoconjugates in context of three-dimensional structures is important in understanding their biol. roles and developing efficient therapeutic agents. Computational modeling and mol. simulation have become essential tool complementary to exptl. methods. Here, we present computational tool, Glycan Modeler for in silico N-/O-glycosylation of target protein and generation of carbohydrate-only systems. In our previous study, we developed Glycan Reader, web-based tool for detecting carbohydrate mols. from PDB structure and generation of simulation system and input files. As integrated into Glycan Reader in CHARMM-GUI, Glycan Modeler enables to generate structures of glycans and glycoconjugates for given glycan sequences and glycosylation sites using PDB glycan template structures from Glycan Fragment Database (http://glycanstructure.org/ fragment-db). Our benchmark tests demonstrate universal applicability of Glycan Reader & Modeler to various glycan sequences and target proteins. We also investigated structural properties of modeled glycan structures by running 2-μs mol. dynamics simulations of HIV envelope protein. Simulations show that modeled glycan structures built by Glycan Reader & Modeler have similar structural features compared to ones solved by X-ray crystallog. We also describe representative examples of glycoconjugate modeling with video demos to illustrate practical applications of Glycan Reader & Modeler.
- 61Jo, S.; Song, K. C.; Desaire, H.; MacKerell, A. D., Jr; Im, W. Glycan Reader: automated sugar identification and simulation preparation for carbohydrates and glycoproteins. J. Comput. Chem. 2011, 32, 3135– 3141, DOI: 10.1002/jcc.2188661https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXpvVamtrk%253D&md5=0cd04c0fb892d58ae69501f23ab079dfGlycan Reader: Automated sugar identification and simulation preparation for carbohydrates and glycoproteinsJo, Sunhwan; Song, Kevin C.; Desaire, Heather; MacKerell, Alexander D.; Im, WonpilJournal of Computational Chemistry (2011), 32 (14), 3135-3141CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)Understanding how glycosylation affects protein structure, dynamics, and function is an emerging and challenging problem in biol. As a first step toward glycan modeling in the context of structural glycobiol., the authors have developed Glycan Reader and integrated it into the CHARMM-GUI,. Glycan Reader greatly simplifies the reading of PDB structure files contg. glycans through (i) detection of carbohydrate mols., (ii) automatic annotation of carbohydrates based on their three-dimensional structures, (iii) recognition of glycosidic linkages between carbohydrates as well as N-/O-glycosidic linkages to proteins, and (iv) generation of inputs for the biomol. simulation program CHARMM with the proper glycosidic linkage setup. In addn., Glycan Reader is linked to other functional modules in CHARMM-GUI, allowing users to easily generate carbohydrate or glycoprotein mol. simulation systems in soln. or membrane environments and visualize the electrostatic potential on glycoprotein surfaces. These tools are useful for studying the impact of glycosylation on protein structure and dynamics. © 2011 Wiley Periodicals, Inc. J Comput Chem, 2011.
- 62Park, S.-J.; Lee, J.; Patel, D. S.; Ma, H.; Lee, H. S.; Jo, S.; Im, W. Glycan Reader is improved to recognize most sugar types and chemical modifications in the Protein Data Bank. Bioinformatics 2017, 33, 3051– 3057, DOI: 10.1093/bioinformatics/btx35862https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhvFGju7bO&md5=16900496c68d5ffeedb08ed0e67e0b94Glycan Reader is improved to recognize most sugar types and chemical modifications in the Protein Data BankPark, Sang-Jun; Lee, Jumin; Patel, Dhilon S.; Ma, Hongjing; Lee, Hui Sun; Jo, Sunhwan; Im, WonpilBioinformatics (2017), 33 (19), 3051-3057CODEN: BOINFP; ISSN:1367-4811. (Oxford University Press)Motivation: Glycans play a central role in many essential biol. processes. Glycan Reader was originally developed to simplify the reading of Protein Data Bank (PDB) files contg. glycans through the automatic detection and annotation of sugars and glycosidic linkages between sugar units and to proteins, all based on at. coordinates and connectivity information. Carbohydrates can have various chem. modifications at different positions, making their chem. space much diverse. Unfortunately, current PDB files do not provide exact annotations for most carbohydrate derivs. and more than 50% of PDB glycan chains have at least one carbohydrate deriv. that could not be correctly recognized by the original Glycan Reader. Results: Glycan Reader has been improved and now identifies most sugar types and chem. modifications (including various glycolipids) in the PDB, and both PDB and PDBx/mmCIF formats are supported. CHARMM-GUI Glycan Reader is updated to generate the simulation system and input of various glycoconjugates with most sugar types and chem. modifications. It also offers a new functionality to edit the glycan structures through addn./deletion/modification of glycosylation types, sugar types, chem. modifications, glycosidic linkages, and anomeric states. The simulation system and input files can be used for CHARMM, NAMD, GROMACS, AMBER, GENESIS, LAMMPS, Desmond, OpenMM, and CHARMM/OpenMM. Glycan Fragment Database in GlycanStructure.Org is also updated to provide an intuitive glycan sequence search tool for complex glycan structures with various chem. modifications in the PDB.
- 63Hanwell, M. D.; Curtis, D. E.; Lonie, D. C.; Vandermeersch, T.; Zurek, E.; Hutchison, G. R. Avogadro: an advanced semantic chemical editor, visualization, and analysis platform. J. Cheminf. 2012, 4, 1– 17, DOI: 10.1186/1758-2946-4-17There is no corresponding record for this reference.
- 64Huang, J.; Rauscher, S.; Nawrocki, G.; Ran, T.; Feig, M.; De Groot, B. L.; Grubmüller, H.; MacKerell, A. D. CHARMM36m: an improved force field for folded and intrinsically disordered proteins. Nat. Methods 2017, 14, 71– 73, DOI: 10.1038/nmeth.406764https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhvVSiu77I&md5=0aa151fbef2ee0b5e2cfb593c54330c2CHARMM36m: an improved force field for folded and intrinsically disordered proteinsHuang, Jing; Rauscher, Sarah; Nawrocki, Grzegorz; Ran, Ting; Feig, Michael; de Groot, Bert L.; Grubmuller, Helmut; MacKerell, Alexander D. JrNature Methods (2017), 14 (1), 71-73CODEN: NMAEA3; ISSN:1548-7091. (Nature Publishing Group)The all-atom additive CHARMM36 protein force field is widely used in mol. modeling and simulations. We present its refinement, CHARMM36m (http://mackerell.umaryland.edu/charmm_ff.shtml), with improved accuracy in generating polypeptide backbone conformational ensembles for intrinsically disordered peptides and proteins.
- 65Guvench, O.; Mallajosyula, S. S.; Raman, E. P.; Hatcher, E.; Vanommeslaeghe, K.; Foster, T. J.; Jamison, F. W.; MacKerell, A. D., Jr CHARMM additive all-atom force field for carbohydrate derivatives and its utility in polysaccharide and carbohydrate-protein modeling. J. Chem. Theory Comput. 2011, 7, 3162– 3180, DOI: 10.1021/ct200328p65https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtFSru77P&md5=fc3c05ef4c95895ee3e7b08472ef3554CHARMM Additive All-Atom Force Field for Carbohydrate Derivatives and Its Utility in Polysaccharide and Carbohydrate-Protein ModelingGuvench, Olgun; Mallajosyula, Sairam S.; Raman, E. Prabhu; Hatcher, Elizabeth; Vanommeslaeghe, Kenno; Foster, Theresa J.; Jamison, Francis W.; MacKerell, Alexander D.Journal of Chemical Theory and Computation (2011), 7 (10), 3162-3180CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Monosaccharide derivs. such as xylose, fucose, N-acetylglucosamine (GlcNAc), N-acetylgalactosamine (GlaNAc), glucuronic acid, iduronic acid, and N-acetylneuraminic acid (Neu5Ac) are important components of eukaryotic glycans. The present work details the development of force-field parameters for these monosaccharides and their covalent connections to proteins via O linkages to serine or threonine side chains and via N linkages to asparagine side chains. The force field development protocol was designed to explicitly yield parameters that are compatible with the existing CHARMM additive force field for proteins, nucleic acids, lipids, carbohydrates, and small mols. Therefore, when combined with previously developed parameters for pyranose and furanose monosaccharides, for glycosidic linkages between monosaccharides, and for proteins, the present set of parameters enables the mol. simulation of a wide variety of biol. important mols. such as complex carbohydrates and glycoproteins. Parametrization included fitting to quantum mech. (QM) geometries and conformational energies of model compds., as well as to QM pair interaction energies and distances of model compds. with water. Parameters were validated in the context of crystals of relevant monosaccharides, as well NMR and/or x-ray crystallog. data on larger systems including oligomeric hyaluronan, sialyl Lewis X, O- and N-linked glycopeptides, and a lectin:sucrose complex. As the validated parameters are an extension of the CHARMM all-atom additive biomol. force field, they further broaden the types of heterogeneous systems accessible with a consistently developed force-field model.
- 66Mallajosyula, S. S.; Guvench, O.; Hatcher, E.; MacKerell, A. D., Jr CHARMM additive all-atom force field for phosphate and sulfate linked to carbohydrates. J. Chem. Theory Comput. 2012, 8, 759– 776, DOI: 10.1021/ct200792v66https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhs1Oju7zP&md5=8350bcdedba6a5e80675f9cae77dfc7eCHARMM Additive All-Atom Force Field for Phosphate and Sulfate Linked to CarbohydratesMallajosyula, Sairam S.; Guvench, Olgun; Hatcher, Elizabeth; MacKerell, Alexander D.Journal of Chemical Theory and Computation (2012), 8 (2), 759-776CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Presented is an extension of the CHARMM additive all-atom carbohydrate force field to enable the modeling of phosphate and sulfate linked to carbohydrates. The parameters are developed in a hierarchical fashion using model compds. contg. the key atoms in the full carbohydrates. Target data for parameter optimization included full two-dimensional energy surfaces defined by the glycosidic dihedral angle pairs in the phosphate/sulfate model compd. analogs of hexopyranose monosaccharide phosphates and sulfates, as detd. by quantum mech. (QM) MP2/cc-pVTZ single point energies on MP2/6-31+G(d) optimized structures. To achieve balanced, transferable dihedral parameters for the dihedral angles, surfaces for all possible anomeric and conformational states were included during the parametrization process. To model physiol. relevant systems, both the mono- and dianionic charged states were studied for the phosphates. This resulted in over 7000 MP2/cc-pVTZ//MP2/6-31G+(d) model compd. conformational energies which, supplemented with QM geometries, were the main target data for the parametrization. Parameters were validated against crystals of relevant monosaccharide derivs. obtained from the Cambridge Structural Database (CSD) and larger systems, inositol-(tri/tetra/penta) phosphates noncovalently bound to the pleckstrin homol. (PH) domain and oligomeric chondroitin sulfate in soln. and in complex with cathepsin K protein.
- 67Vanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.; Mackerell, A. D. CHARMM general force field: A force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fields. J. Comput. Chem. 2010, 31, 671– 690, DOI: 10.1002/jcc.2136767https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhtlentbc%253D&md5=26212e0e4f73bded0c89d4b411cd3833CHARMM general force field: A force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fieldsVanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.; Mackerell, A. D., Jr.Journal of Computational Chemistry (2010), 31 (4), 671-690CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)The widely used CHARMM additive all-atom force field includes parameters for proteins, nucleic acids, lipids, and carbohydrates. In the present article, an extension of the CHARMM force field to drug-like mols. is presented. The resulting CHARMM General Force Field (CGenFF) covers a wide range of chem. groups present in biomols. and drug-like mols., including a large no. of heterocyclic scaffolds. The parametrization philosophy behind the force field focuses on quality at the expense of transferability, with the implementation concg. on an extensible force field. Statistics related to the quality of the parametrization with a focus on exptl. validation are presented. Addnl., the parametrization procedure, described fully in the present article in the context of the model systems, pyrrolidine, and 3-phenoxymethyl-pyrrolidine will allow users to readily extend the force field to chem. groups that are not explicitly covered in the force field as well as add functional groups to and link together mols. already available in the force field. CGenFF thus makes it possible to perform "all-CHARMM" simulations on drug-target interactions thereby extending the utility of CHARMM force fields to medicinally relevant systems. © 2009 Wiley Periodicals, Inc.J Comput Chem, 2010.
- 68Yu, W.; He, X.; Vanommeslaeghe, K.; MacKerell, A. D., Jr Extension of the CHARMM general force field to sulfonyl-containing compounds and its utility in biomolecular simulations. J. Comput. Chem. 2012, 33, 2451– 2468, DOI: 10.1002/jcc.2306768https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhtVKntL7J&md5=8de6f9a92b8ad706b010534a51fe54cdExtension of the CHARMM general force field to sulfonyl-containing compounds and its utility in biomolecular simulationsYu, Wenbo; He, Xibing; Vanommeslaeghe, Kenno; MacKerell, Alexander D.Journal of Computational Chemistry (2012), 33 (31), 2451-2468CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)Presented is an extension of the CHARMM General Force Field (CGenFF) to enable the modeling of sulfonyl-contg. compds. Model compds. contg. chem. moieties such as sulfone, sulfonamide, sulfonate, and sulfamate were used as the basis for the parameter optimization. Targeting high-level quantum mech. and exptl. crystal data, the new parameters were optimized in a hierarchical fashion designed to maintain compatibility with the remainder of the CHARMM additive force field. The optimized parameters satisfactorily reproduced equil. geometries, vibrational frequencies, interactions with water, gas phase dipole moments, and dihedral potential energy scans. Validation involved both cryst. and liq. phase calcns. showing the newly developed parameters to satisfactorily reproduce exptl. unit cell geometries, crystal intramol. geometries, and pure solvent densities. The force field was subsequently applied to study conformational preference of a sulfonamide-based peptide system. Good agreement with exptl. IR/NMR data further validated the newly developed CGenFF parameters as a tool to investigate the dynamic behavior of sulfonyl groups in a biol. environment. CGenFF now covers sulfonyl group contg. moieties allowing for modeling and simulation of sulfonyl-contg. compds. in the context of biomol. systems including compds. of medicinal interest.
- 69Vanommeslaeghe, K.; MacKerell, A. D., Jr Automation of the CHARMM General Force Field (CGenFF) I: bond perception and atom typing. J. Chem. Inf. Model. 2012, 52, 3144– 3154, DOI: 10.1021/ci300363c69https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhs1Gns7fL&md5=c6679293f4a2501f2bcadf2020ca1473Automation of the CHARMM General Force Field (CGenFF) I: Bond Perception and Atom TypingVanommeslaeghe, K.; MacKerell, A. D.Journal of Chemical Information and Modeling (2012), 52 (12), 3144-3154CODEN: JCISD8; ISSN:1549-9596. (American Chemical Society)Mol. mechanics force fields are widely used in computer-aided drug design for the study of drug-like mols. alone or interacting with biol. systems. In simulations involving biol. macromols., the biol. part is typically represented by a specialized biomol. force field, while the drug is represented by a matching general (org.) force field. In order to apply these general force fields to an arbitrary drug-like mol., functionality for assignment of atom types, parameters, and charges is required. In the present article, which is part I of a series of two, we present the algorithms for bond perception and atom typing for the CHARMM General Force Field (CGenFF). The CGenFF atom typer first assocs. attributes to the atoms and bonds in a mol., such as valence, bond order, and ring membership among others. Of note are a no. of features that are specifically required for CGenFF. This information is then used by the atom typing routine to assign CGenFF atom types based on a programmable decision tree. This allows for straight-forward implementation of CGenFF's complicated atom typing rules and for equally straight-forward updating of the atom typing scheme as the force field grows. The presented atom typer was validated by assigning correct atom types on 477 model compds. including in the training set as well as 126 test-set mols. that were constructed to specifically verify its different components. The program may be utilized via an online implementation at https://www.paramchem.org/.
- 70Vanommeslaeghe, K.; Raman, E. P.; MacKerell, A. D., Jr Automation of the CHARMM General Force Field (CGenFF) II: assignment of bonded parameters and partial atomic charges. J. Chem. Inf. Model. 2012, 52, 3155– 3168, DOI: 10.1021/ci300364970https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhs1Gns7fF&md5=e676ad1f42cb1e98dd353d4d285e8d13Automation of the CHARMM General Force Field (CGenFF) II: Assignment of Bonded Parameters and Partial Atomic ChargesVanommeslaeghe, K.; Raman, E. Prabhu; MacKerell, A. D.Journal of Chemical Information and Modeling (2012), 52 (12), 3155-3168CODEN: JCISD8; ISSN:1549-9596. (American Chemical Society)Mol. mechanics force fields are widely used in computer-aided drug design for the study of drug candidates interacting with biol. systems. In these simulations, the biol. part is typically represented by a specialized biomol. force field, while the drug is represented by a matching general (org.) force field. In order to apply these general force fields to an arbitrary drug-like mol., functionality for assignment of atom types, parameters, and partial at. charges is required. In the present article, algorithms for the assignment of parameters and charges for the CHARMM General Force Field (CGenFF) are presented. These algorithms rely on the existing parameters and charges that were detd. as part of the parametrization of the force field. Bonded parameters are assigned based on the similarity between the atom types that define said parameters, while charges are detd. using an extended bond-charge increment scheme. Charge increments were optimized to reproduce the charges on model compds. that were part of the parametrization of the force field. Case studies are presented to clarify the functioning of the algorithms and the significance of their output data.
- 71Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R. W.; Klein, M. L. Comparison of simple potential functions for simulating liquid water. J. Chem. Phys. 1983, 79, 926– 935, DOI: 10.1063/1.44586971https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3sXksF2htL4%253D&md5=a1161334e381746be8c9b15a5e56f704Comparison of simple potential functions for simulating liquid waterJorgensen, William L.; Chandrasekhar, Jayaraman; Madura, Jeffry D.; Impey, Roger W.; Klein, Michael L.Journal of Chemical Physics (1983), 79 (2), 926-35CODEN: JCPSA6; ISSN:0021-9606.Classical Monte Carlo simulations were carried out for liq. H2O in the NPT ensemble at 25° and 1 atm using 6 of the simpler intermol. potential functions for the dimer. Comparisons were made with exptl. thermodn. and structural data including the neutron diffraction results of Thiessen and Narten (1982). The computed densities and potential energies agree with expt. except for the original Bernal-Fowler model, which yields an 18% overest. of the d. and poor structural results. The discrepancy may be due to the correction terms needed in processing the neutron data or to an effect uniformly neglected in the computations. Comparisons were made for the self-diffusion coeffs. obtained from mol. dynamics simulations.
- 72MacKerell, A. D., Jr; Bashford, D.; Bellott, M.; Dunbrack, R. L., Jr; Evanseck, J. D.; Field, M. J.; Fischer, S.; Gao, J.; Guo, H.; Ha, S. All-atom empirical potential for molecular modeling and dynamics studies of proteins. J. Phys. Chem. B 1998, 102, 3586– 3616, DOI: 10.1021/jp973084f72https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXivVOlsb4%253D&md5=ebb5100dafd0daeee60ca2fa66c1324aAll-Atom Empirical Potential for Molecular Modeling and Dynamics Studies of ProteinsMacKerell, A. D., Jr.; Bashford, D.; Bellott, M.; Dunbrack, R. L.; Evanseck, J. D.; Field, M. J.; Fischer, S.; Gao, J.; Guo, H.; Ha, S.; Joseph-McCarthy, D.; Kuchnir, L.; Kuczera, K.; Lau, F. T. K.; Mattos, C.; Michnick, S.; Ngo, T.; Nguyen, D. T.; Prodhom, B.; Reiher, W. E., III; Roux, B.; Schlenkrich, M.; Smith, J. C.; Stote, R.; Straub, J.; Watanabe, M.; Wiorkiewicz-Kuczera, J.; Yin, D.; Karplus, M.Journal of Physical Chemistry B (1998), 102 (18), 3586-3616CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)New protein parameters are reported for the all-atom empirical energy function in the CHARMM program. The parameter evaluation was based on a self-consistent approach designed to achieve a balance between the internal (bonding) and interaction (nonbonding) terms of the force field and among the solvent-solvent, solvent-solute, and solute-solute interactions. Optimization of the internal parameters used exptl. gas-phase geometries, vibrational spectra, and torsional energy surfaces supplemented with ab initio results. The peptide backbone bonding parameters were optimized with respect to data for N-methylacetamide and the alanine dipeptide. The interaction parameters, particularly the at. charges, were detd. by fitting ab initio interaction energies and geometries of complexes between water and model compds. that represented the backbone and the various side chains. In addn., dipole moments, exptl. heats and free energies of vaporization, solvation and sublimation, mol. vols., and crystal pressures and structures were used in the optimization. The resulting protein parameters were tested by applying them to noncyclic tripeptide crystals, cyclic peptide crystals, and the proteins crambin, bovine pancreatic trypsin inhibitor, and carbonmonoxy myoglobin in vacuo and in a crystal. A detailed anal. of the relationship between the alanine dipeptide potential energy surface and calcd. protein φ, χ angles was made and used in optimizing the peptide group torsional parameters. The results demonstrate that use of ab initio structural and energetic data by themselves are not sufficient to obtain an adequate backbone representation for peptides and proteins in soln. and in crystals. Extensive comparisons between mol. dynamics simulation and exptl. data for polypeptides and proteins were performed for both structural and dynamic properties. Calcd. data from energy minimization and dynamics simulations for crystals demonstrate that the latter are needed to obtain meaningful comparisons with exptl. crystal structures. The presented parameters, in combination with the previously published CHARMM all-atom parameters for nucleic acids and lipids, provide a consistent set for condensed-phase simulations of a wide variety of mols. of biol. interest.
- 73Venable, R. M.; Luo, Y.; Gawrisch, K.; Roux, B.; Pastor, R. W. Simulations of anionic lipid membranes: development of interaction-specific ion parameters and validation using NMR data. J. Phys. Chem. B 2013, 117, 10183– 10192, DOI: 10.1021/jp401512z73https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXht1ars7%252FL&md5=f479595c0149e02b4b9e663b772120b9Simulations of Anionic Lipid Membranes: Development of Interaction-Specific Ion Parameters and Validation Using NMR DataVenable, Richard M.; Luo, Yun; Gawrisch, Klaus; Roux, Benoit; Pastor, Richard W.Journal of Physical Chemistry B (2013), 117 (35), 10183-10192CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)Overbinding of ions to lipid head groups is a potentially serious artifact in simulations of charged lipid bilayers. In this study, the Lennard-Jones radii in the CHARMM force field for interactions of Na+ and lipid oxygen atoms of carboxyl, phosphate, and ester groups were revised to match osmotic pressure data on sodium acetate and electrophoresis data on palmitoyloleoyl phosphatidylcholine (POPC) vesicles. The new parameters were then validated by successfully reproducing previously published exptl. NMR deuterium order parameters for dimyristoyl phosphatidylglycerol (DMPG) and newly obtained values for palmitoyloleoyl phosphatidylserine (POPS). Although the increases in Lennard-Jones diams. are only 0.02-0.12 Å, they are sufficient to reduce Na+ binding, and thereby increase surface areas per lipid by 5-10% compared with the unmodified parameters.
- 74Abraham, M. J.; Murtola, T.; Schulz, R.; Páll, S.; Smith, J. C.; Hess, B.; Lindahl, E. GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 2015, 1-2, 19– 25, DOI: 10.1016/j.softx.2015.06.001There is no corresponding record for this reference.
- 75Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling through velocity rescaling. J. Chem. Phys. 2007, 126, 014101 DOI: 10.1063/1.240842075https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXosVCltg%253D%253D&md5=9c182b57bfc65bca6be23c8c76b4be77Canonical sampling through velocity rescalingBussi, Giovanni; Donadio, Davide; Parrinello, MicheleJournal of Chemical Physics (2007), 126 (1), 014101/1-014101/7CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The authors present a new mol. dynamics algorithm for sampling the canonical distribution. In this approach the velocities of all the particles are rescaled by a properly chosen random factor. The algorithm is formally justified and it is shown that, in spite of its stochastic nature, a quantity can still be defined that remains const. during the evolution. In numerical applications this quantity can be used to measure the accuracy of the sampling. The authors illustrate the properties of this new method on Lennard-Jones and TIP4P water models in the solid and liq. phases. Its performance is excellent and largely independent of the thermostat parameter also with regard to the dynamic properties.
- 76Parrinello, M.; Rahman, A. Polymorphic transitions in single crystals: A new molecular dynamics method. J. Appl. Phys. 1981, 52, 7182– 7190, DOI: 10.1063/1.32869376https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL38XislSnuw%253D%253D&md5=a0a5617389f6cabbf2a405c649aadf03Polymorphic transitions in single crystals: A new molecular dynamics methodParrinello, M.; Rahman, A.Journal of Applied Physics (1981), 52 (12), 7182-90CODEN: JAPIAU; ISSN:0021-8979.A Lagrangian formulation is introduced; it can be used to make mol. dynamics (MD) calcns. on systems under the most general, externally applied, conditions of stress. In this formulation the MD cell shape and size can change according to dynamic equations given by this Lagrangian. This MD technique was used to the study of structural transitions of a Ni single crystal under uniform uniaxial compressive and tensile loads. Some results regarding the stress-strain relation obtained by static calcns. are invalid at finite temp. Under compressive loading, the model of Ni shows a bifurcation in its stress-strain relation; this bifurcation provides a link in configuration space between cubic and hexagonal close packing. Such a transition could perhaps be obsd. exptl. under extreme conditions of shock.
- 77Hess, B. P-LINCS: A A parallel linear constraint solver for molecular simulation. J. Chem. Theory Comput. 2008, 4, 116– 122, DOI: 10.1021/ct700200b77https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXhtlKru7zL&md5=476d5ca2eb25574d44b775996fff7b75P-LINCS: A Parallel Linear Constraint Solver for Molecular SimulationHess, BerkJournal of Chemical Theory and Computation (2008), 4 (1), 116-122CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)By removing the fastest degrees of freedom, constraints allow for an increase of the time step in mol. simulations. In the last decade parallel simulations have become commonplace. However, up till now efficient parallel constraint algorithms have not been used with domain decompn. In this paper the parallel linear constraint solver (P-LINCS) is presented, which allows the constraining of all bonds in macromols. Addnl. the energy conservation properties of (P-)LINCS are assessed in view of improvements in the accuracy of uncoupled angle constraints and integration in single precision.
- 78Darden, T.; York, D.; Pedersen, L. Particle mesh Ewald: An N · log(N) method for Ewald sums in large systems. J. Chem. Phys. 1993, 98, 10089– 10092, DOI: 10.1063/1.46439778https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXks1Ohsr0%253D&md5=3c9f230bd01b7b714fd096d4d2e755f6Particle mesh Ewald: an N·log(N) method for Ewald sums in large systemsDarden, Tom; York, Darrin; Pedersen, LeeJournal of Chemical Physics (1993), 98 (12), 10089-92CODEN: JCPSA6; ISSN:0021-9606.An N·log(N) method for evaluating electrostatic energies and forces of large periodic systems is presented. The method is based on interpolation of the reciprocal space Ewald sums and evaluation of the resulting convolution using fast Fourier transforms. Timings and accuracies are presented for three large cryst. ionic systems.
- 79Loche, P.; Steinbrunner, P.; Friedowitz, S.; Netz, R. R.; Bonthuis, D. J. Transferable Ion Force Fields in Water from a Simultaneous Optimization of Ion Solvation and Ion–Ion Interaction. J. Phys. Chem. B 2021, 125, 8581– 8587, DOI: 10.1021/acs.jpcb.1c0530379https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhsF2lur3L&md5=f632bce7d53b0fc8f8d8763205915a02Transferable Ion Force Fields in Water from a Simultaneous Optimization of Ion Solvation and Ion-Ion InteractionLoche, Philip; Steinbrunner, Patrick; Friedowitz, Sean; Netz, Roland R.; Bonthuis, Douwe JanJournal of Physical Chemistry B (2021), 125 (30), 8581-8587CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)The poor performance of many existing nonpolarizable ion force fields is typically blamed on either the lack of explicit polarizability, the absence of charge transfer, or the use of unreduced Coulomb interactions. However, this anal. disregards the large and mostly unexplored parameter range offered by the Lennard-Jones potential. We use a global optimization procedure to develop water-model-transferable force fields for the ions K+, Na+, Cl-, and Br- in the complete parameter space of all Lennard-Jones interactions using std. mixing rules. No extra-thermodn. assumption is necessary for the simultaneous optimization of the four ion pairs. After optimization with respect to the exptl. solvation free energy and activity, the force fields reproduce the concn. dependent d., ionic cond. and dielec. const. with high accuracy. The force field is fully transferable between SPC/E, TIP3P, and TIP4P/ε water models. Our results show that a thermodynamically consistent force field for these ions needs only Lennard-Jones and std. Coulomb interactions.
- 80Humphrey, W.; Dalke, A.; Schulten, K. VMD: visual molecular dynamics. J. Mol. Graphics 1996, 14, 33– 38, DOI: 10.1016/0263-7855(96)00018-580https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28Xis12nsrg%253D&md5=1e3094ec3151fb85c5ff05f8505c78d5VDM: visual molecular dynamicsHumphrey, William; Dalke, Andrew; Schulten, KlausJournal of Molecular Graphics (1996), 14 (1), 33-8, plates, 27-28CODEN: JMGRDV; ISSN:0263-7855. (Elsevier)VMD is a mol. graphics program designed for the display and anal. of mol. assemblies, in particular, biopolymers such as proteins and nucleic acids. VMD can simultaneously display any no. of structures using a wide variety of rendering styles and coloring methods. Mols. are displayed as one or more "representations," in which each representation embodies a particular rendering method and coloring scheme for a selected subset of atoms. The atoms displayed in each representation are chosen using an extensive atom selection syntax, which includes Boolean operators and regular expressions. VMD provides a complete graphical user interface for program control, as well as a text interface using the Tcl embeddable parser to allow for complex scripts with variable substitution, control loops, and function calls. Full session logging is supported, which produces a VMD command script for later playback. High-resoln. raster images of displayed mols. may be produced by generating input scripts for use by a no. of photorealistic image-rendering applications. VMD has also been expressly designed with the ability to animate mol. dynamics (MD) simulation trajectories, imported either from files or from a direct connection to a running MD simulation. VMD is the visualization component of MDScope, a set of tools for interactive problem solving in structural biol., which also includes the parallel MD program NAMD, and the MDCOMM software used to connect the visualization and simulation programs, VMD is written in C++, using an object-oriented design; the program, including source code and extensive documentation, is freely available via anonymous ftp and through the World Wide Web.
- 81Gowers, R. J.; Linke, M.; Barnoud, J.; Reddy, T. J.; Melo, M. N.; Seyler, S. L.; Domanski, J.; Dotson, D. L.; Buchoux, S.; Kenney, I. M.; Beckstein, O. MDAnalysis: A Python Package for the Rapid Analysis of Molecular Dynamics Simulations; Proceedings of the 15th Python in Science Conference; Office of Scientific and Technical Information: Los Alamos, NM, 2016; p 105.There is no corresponding record for this reference.
- 82Michaud-Agrawal, N.; Denning, E. J.; Woolf, T. B.; Beckstein, O. MDAnalysis: a toolkit for the analysis of molecular dynamics simulations. J. Comput. Chem. 2011, 32, 2319– 2327, DOI: 10.1002/jcc.2178782https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXnvFalsr8%253D&md5=d567042c65cfdc1c81336a29137654bfMDAnalysis: A toolkit for the analysis of molecular dynamics simulationsMichaud-Agrawal, Naveen; Denning, Elizabeth J.; Woolf, Thomas B.; Beckstein, OliverJournal of Computational Chemistry (2011), 32 (10), 2319-2327CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)MDAnal. is an object-oriented library for structural and temporal anal. of mol. dynamics (MD) simulation trajectories and individual protein structures. It is written in the Python language with some performance-crit. code in C. It uses the powerful NumPy package to expose trajectory data as fast and efficient NumPy arrays. It has been tested on systems of millions of particles. Many common file formats of simulation packages including CHARMM, Gromacs, Amber, and NAMD and the Protein Data Bank format can be read and written. Atoms can be selected with a syntax similar to CHARMM's powerful selection commands. MDAnal. enables both novice and experienced programmers to rapidly write their own anal. tools and access data stored in trajectories in an easily accessible manner that facilitates interactive explorative anal. MDAnal. has been tested on and works for most Unix-based platforms such as Linux and Mac OS X. It is freely available under the GNU General Public License from http://mdanal.googlecode.com. © 2011 Wiley Periodicals, Inc. J Comput Chem 2011.
- 83Schwarzl, R.; Liese, S.; Brünig, F. N.; Laudisio, F.; Netz, R. R. Force Response of Polypeptide Chains from Water-Explicit MD Simulations. Macromolecules 2020, 53, 4618– 4629, DOI: 10.1021/acs.macromol.0c0013883https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhtFeltrnK&md5=0592929f6ff286df1de443f5f0c3748fForce Response of Polypeptide Chains from Water-Explicit MD SimulationsSchwarzl, Richard; Liese, Susanne; Bruenig, Florian N.; Laudisio, Fabio; Netz, Roland R.Macromolecules (Washington, DC, United States) (2020), 53 (12), 4618-4629CODEN: MAMOBX; ISSN:0024-9297. (American Chemical Society)Using mol. dynamics simulations in explicit water, the force-extension relations for the five homopeptides polyglycine, polyalanine, polyasparagine, poly(glutamic acid), and polylysine are investigated. From simulations in the low-force regime the Kuhn length is detd., from simulations in the high-force regime the equil. contour length and the linear and nonlinear stretching moduli, which agree well with quantum-chem. d.-functional theory calcns., are detd. All these parameters vary considerably between the different polypeptides. The augmented inhomogeneous partially freely rotating chain (iPFRC) model, which accounts for side-chain interactions and restricted dihedral rotation, is demonstrated to describe the simulated force-extension relations very well. We present a quant. comparison between published exptl. single-mol. force-extension curves for different polypeptides with simulation and model predictions. The thermodn. stretching properties of polypeptides are investigated by decompn. of the stretching free energy into energetic and entropic contributions.
- 84Flyvbjerg, H.; Petersen, H. G. Error estimates on averages of correlated data. J. Chem. Phys. 1989, 91, 461– 466, DOI: 10.1063/1.45748084https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXkslOis7s%253D&md5=a424bd42502aa2989d102aa4c87b8ecaError estimates on averages of correlated dataFlyvbjerg, H.; Petersen, H. G.Journal of Chemical Physics (1989), 91 (1), 461-6CODEN: JCPSA6; ISSN:0021-9606.A description is given on how the true statistical error on an av. of correlated data can be obtained with ease and efficiency by a renormalization group method. The method is illustrated with numerical and anal. examples, having finite as well as infinite range correlations.
Supporting Information
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacsau.4c00886.
Derivation of the standard free-energy of binding from a polymer desorption free-energy profile; additional figures (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.