Hostname: page-component-76fb5796d-25wd4 Total loading time: 0 Render date: 2024-04-30T06:06:27.080Z Has data issue: false hasContentIssue false

Towards design of drugs and delivery systems with the Martini coarse-grained model

Published online by Cambridge University Press:  12 October 2022

Lisbeth R. Kjølbye
Affiliation:
Molecular Microbiology and Structural Biochemistry (MMSB, UMR 5086), CNRS & University of Lyon, Lyon, France
Gilberto P. Pereira
Affiliation:
Molecular Microbiology and Structural Biochemistry (MMSB, UMR 5086), CNRS & University of Lyon, Lyon, France
Alessio Bartocci
Affiliation:
Institut de Chimie de Strasbourg, UMR 7177 CNRS, Université de Strasbourg, Strasbourg Cedex, France
Martina Pannuzzo
Affiliation:
PharmCADD, Busan, South Korea
Simone Albani
Affiliation:
Computational Biomedicine, Institute of Advanced Simulation (IAS-5) and Institute of Neuroscience and Medicine (INM-9), Forschungszentrum Jülich GmbH, Jülich, Germany Department of Biology, Faculty of Mathematics, Computer Science and Natural Sciences, RWTH Aachen University, Aachen, Germany
Alessandro Marchetto
Affiliation:
Computational Biomedicine, Institute of Advanced Simulation (IAS-5) and Institute of Neuroscience and Medicine (INM-9), Forschungszentrum Jülich GmbH, Jülich, Germany Department of Biology, Faculty of Mathematics, Computer Science and Natural Sciences, RWTH Aachen University, Aachen, Germany
Brian Jiménez-García
Affiliation:
Zymvol Biomodeling, Barcelona, Spain
Juliette Martin
Affiliation:
Molecular Microbiology and Structural Biochemistry (MMSB, UMR 5086), CNRS & University of Lyon, Lyon, France
Giulia Rossetti
Affiliation:
Computational Biomedicine, Institute of Advanced Simulation (IAS-5) and Institute of Neuroscience and Medicine (INM-9), Forschungszentrum Jülich GmbH, Jülich, Germany Jülich Supercomputing Centre (JSC), Forschungszentrum Jülich GmbH, Jülich, Germany Department of Neurology, Faculty of Medicine, RWTH Aachen University, Aachen, Germany
Marco Cecchini
Affiliation:
Institut de Chimie de Strasbourg, UMR 7177 CNRS, Université de Strasbourg, Strasbourg Cedex, France
Sangwook Wu
Affiliation:
PharmCADD, Busan, South Korea Department of Physics, Pukyong National University, Busan, Republic of Korea
Luca Monticelli
Affiliation:
Molecular Microbiology and Structural Biochemistry (MMSB, UMR 5086), CNRS & University of Lyon, Lyon, France
Paulo C. T. Souza*
Affiliation:
Molecular Microbiology and Structural Biochemistry (MMSB, UMR 5086), CNRS & University of Lyon, Lyon, France
*
*Author for correspondence: Paulo C. T. Souza, E-mail: paulo.telles-de-souza@ibcp.fr
Rights & Permissions [Opens in a new window]

Abstract

Coarse-grained (CG) modelling with the Martini force field has come of age. By combining a variety of bead types and sizes with a new mapping approach, the newest version of the model is able to accurately simulate large biomolecular complexes at millisecond timescales. In this perspective, we discuss possible applications of the Martini 3 model in drug discovery and development pipelines and highlight areas for future development. Owing to its high simulation efficiency and extended chemical space, Martini 3 has great potential in the area of drug design and delivery. However, several aspects of the model should be improved before Martini 3 CG simulations can be routinely employed in academic and industrial settings. These include the development of automatic parameterisation protocols for a variety of molecule types, the improvement of backmapping procedures, the description of protein flexibility and the development of methodologies enabling efficient sampling. We illustrate our view with examples on key areas where Martini could give important contributions such as drugs targeting membrane proteins, cryptic pockets and protein–protein interactions and the development of soft drug delivery systems.

Type
Perspective
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2022. Published by Cambridge University Press

Introduction

Recent studies have shown that the cost of drug discovery and development is, on average, higher than several hundred million dollars (Mohs and Greig, Reference Mohs and Greig2017; Schlander et al., Reference Schlander, Hernandez-Villafuerte, Cheng, Mestre-Ferrandiz and Baumann2021). Moreover, several diseases such as Alzheimer, cancer, viral infections and cardiovascular diseases remain orphan of an effective, long-term and safe therapeutic protocol (Falzone et al., Reference Falzone, Salomone and Libra2018; Nishiga et al., Reference Nishiga, Wang, Han, Lewis and Wu2020; Brown and Wobst, Reference Brown and Wobst2021; Esang and Gupta, Reference Esang and Gupta2021). Current challenges in the development of novel therapeutic approaches include the unavailability of druggable binding pockets in the target structure (Weerakoon et al., Reference Weerakoon, Carbajo, De Maria, Tyrchan and Zhao2022) and the lack of effective delivery systems, which can improve drug pharmacodynamics (Wang et al., Reference Wang, Ye, Gao and Ouyang2021b).

Computational methodologies can speed up the drug discovery pipeline, decrease the associated costs and provide insight into the interactions between drugs and their targets, which is critical for rational drug design (Sliwoski et al., Reference Sliwoski, Kothiwale, Meiler and Lowe2014; Lin et al., Reference Lin, Li and Lin2020). Computer modelling permeates both hit-identification and lead-optimisation stages of drug discovery pipelines. Computational methods have been used to predict protein–ligand binding modes (Śledź and Caflisch, Reference Śledź and Caflisch2018), binding affinities (Montalvo-Acosta and Cecchini, Reference Montalvo-Acosta and Cecchini2016), brain–blood barrier permeation (Crivori et al., Reference Crivori, Cruciani, Carrupt and Testa2000), compound activity against a given target (Pereira et al., Reference Pereira, Szwarc, Mondragão, Lima and Pereira2018) or to identify and map potential binding sites (Yu and MacKerell, Reference Yu and MacKerell2017; MacKerell et al., Reference MacKerell, Jo, Lakkaraju, Lind and Yu2020). Some of these methods rely on atomistic molecular dynamics (MD) simulations to produce configurational ensembles (Siebenmorgen and Zacharias, Reference Siebenmorgen and Zacharias2020). However, converging on sampling the potential energy landscape of large biomolecular complexes is challenging and limits the application of atomistic MD to smaller systems. Nonetheless, numerical simulations and docking studies can still contribute to studies of protein–ligand interactions or identification of hit compounds (Jorgensen, Reference Jorgensen2009; Bollini et al., Reference Bollini, Domaoal, Thakur, Gallardo-Macias, Spasov, Anderson and Jorgensen2011; Frey et al., Reference Frey, Gray, Spasov, Bollini, Gallardo-Macias, Jorgensen and Anderson2013). Alternatively, purpose-built hardware and software can help simulate large systems at atomistic resolution as shown by the DESRES team (Dror et al., Reference Dror, Young, Shaw and Padua2011; Shaw et al., Reference Shaw, Adams, Azaria, Bank, Batson, Bell, Bergdorf, Bhatt, Butts, Correia, Dirks, Dror, Eastwood, Edwards, Even, Feldmann, Fenn, Fenton, Forte, Gagliardo, Gill, Gorlatova, Greskamp, Grossman, Gullingsrud, Harper, Hasenplaugh, Heily, Heshmat, Hunt, Ierardi, Iserovich, Jackson, Johnson, Kirk, Klepeis, Kuskin, Mackenzie, Mader, McGowen, McLaughlin, Moraes, Nasr, Nociolo, O’Donnell, Parker, Peticolas, Pocina, Predescu, Quan, Salmon, Schwink, Shim, Siddique, Spengler, Szalay, Tabladillo, Tartler, Taube, Theobald, Towles, Vick, Wang, Wazlowski, Weingarten, Williams and Yuh2021).

Drug delivery has also seen an increase in usage of computational modelling, mainly because current development pipelines rest upon unpredictable trial and error experiments. Molecular modelling offers an attractive platform for understanding and optimising delivery systems in a biologically relevant context (Wang et al., Reference Wang, Ye, Gao and Ouyang2021b). In this field, the limitations associated with system size and complexity are magnified. Several model systems exploring interaction with lipid bilayers have been constructed, with more realistic models mainly being of solid nanoparticles (NPs) such as gold NPs (Franco-Ulloa et al., Reference Franco-Ulloa, Guarnieri, Riccardi, Pompa and De Vivo2021; Salassi et al., Reference Salassi, Caselli, Cardellini, Lavagna, Montis, Berti and Rossi2021). Limited studies have explored softer delivery systems like lipid-based NPs, mainly due to the lack of well-established computational protocols for constructing and studying these systems.

Coarse-grained (CG) modelling techniques alleviate sampling limitations of atomistic MD. The most widely used CG force field (FF) is the Martini FF (Marrink et al., Reference Marrink, Risselada, Yefimov, Tieleman and De Vries2007). The newly developed Martini 3 (Souza et al., Reference Souza, Alessandri, Barnoud, Thallmair, Faustino, Grünewald, Patmanidis, Abdizadeh, Bruininks, Wassenaar, Kroon, Melcr, Nieto, Corradi, Khan, Domański, Javanainen, Martinez-Seara, Reuter, Best, Vattulainen, Monticelli, Periole, Tieleman, de Vries and Marrink2021a) improves sampling efficiency by merging together two to four non-hydrogen atoms and corresponding associated hydrogens into one interaction bead, with the bonded and non-bonded parameters derived from a combination of bottom-up and top-down approaches, respectively. In parallel with the development of the Martini 3 FF, other CG approaches were pursued. For instance, some recent developments in protein CG models include the SIRAH2.0 FF (Machado et al., Reference Machado, Barrera, Klein, Soñora, Silva and Pantano2019), SPICA (Kawamoto et al., Reference Kawamoto, Liu, Miyazaki, Seo, Dixit, DeVane, MacDermaid, Fiorin, Klein and Shinoda2022) or the recently developed ProMPT, an alternative polarisable Martini model (Sahoo et al., Reference Sahoo, Lee and Matysiak2022).

The Martini 2 FF currently supports a wide array of parameters for proteins, different lipid types, polymers, DNA and RNA (Monticelli et al., Reference Monticelli, Kandasamy, Periole, Larson, Tieleman and Marrink2008; López et al., Reference López, Rzepiela, de Vries, Dijkhuizen, Hünenberger and Marrink2009; De Jong et al., Reference De Jong, Singh, Bennett, Arnarez, Wassenaar, Schäfer, Periole, Tieleman and Marrink2013; Uusitalo et al., Reference Uusitalo, Ingólfsson, Akhshi, Tieleman and Marrink2015, Reference Uusitalo, Ingólfsson, Marrink and Faustino2017; Grünewald et al., Reference Grünewald, Rossi, de Vries, Marrink and Monticelli2018; Salassi et al., Reference Salassi, Simonelli, Bartocci and Rossi2018). Four main bead types were developed based on the polarity of chemical groups. These particles are further subdivided depending, for example, on their hydrogen-bonding capabilities (Marrink et al., Reference Marrink, Risselada, Yefimov, Tieleman and De Vries2007). Limitations of the Martini 2 model included overstabilisation of some biomolecular interactions, mainly noted for proteins and sugars (Alessandri et al., Reference Alessandri, Souza, Thallmair, Melo, De Vries and Marrink2019) and the narrow range of chemical groups represented by the available beads (Kanekal and Bereau, Reference Kanekal and Bereau2019). The new version 3 (Souza et al., Reference Souza, Alessandri, Barnoud, Thallmair, Faustino, Grünewald, Patmanidis, Abdizadeh, Bruininks, Wassenaar, Kroon, Melcr, Nieto, Corradi, Khan, Domański, Javanainen, Martinez-Seara, Reuter, Best, Vattulainen, Monticelli, Periole, Tieleman, de Vries and Marrink2021a) addressed these issues and now provides promising solutions for drug design and delivery. New Martini 3 CG models allow simulations of more complex systems, facilitating the study of important biomolecular processes like ligand binding (Souza et al., Reference Souza, Thallmair, Conflitti, Ramírez-Palacios, Alessandri, Raniolo, Limongelli and Marrink2020), fusion events (Bruininks et al., Reference Bruininks, Souza, Ingólfsson and Marrink2020), and the distribution of drugs within particle or carrier delivery systems (Casalini, Reference Casalini2021). This enables understanding of the forces behind encapsulation and drug release, which furthers the optimisation and development of delivery systems, as well as the interactions, which drive ligand binding, fundamental for drug design campaigns.

The Martini 2 and 3 FFs have been applied to study different biomolecular systems, among them proteins, membranes or vesicles, and is increasingly being used in the field of materials sciences (Marrink et al., Reference Marrink, Corradi, Souza, Ingólfsson, Tieleman and Sansom2019; Alessandri et al., Reference Alessandri, Grünewald and Marrink2021; Marrink et al., Reference Marrink, Monticelli, Melo, Alessandri, Tieleman and Souza2022). Examples exist of CG simulations studying fusion of delivery systems, such as lipoplexes or nanoemulsions, with lipid bilayers (Lee et al., Reference Lee, Schlesinger, Wickline, Lanza and Baker2012; Bruininks et al., Reference Bruininks, Souza, Ingólfsson and Marrink2020; Gupta et al., Reference Gupta, Das and Chow2021; Machado et al., Reference Machado, Bruininks, Singh, Santos, Pizzol, Dieamant, Kruger, Martin, Marrink, Souza and Favero2022). In 2020, Bruininks et al. (Reference Bruininks, Souza, Ingólfsson and Marrink2020) used CG modelling to simulate the fusion of a cationic lipoplex containing DNA with a simple membrane model representing the endosomal membrane. This is one of the first stepping stones for using CG models to explore nucleic acid (NA) release. For drug binding, the potential of CG-Martini simulations in studies of protein–ligand binding was shown in the work of Negami et al. (Reference Negami, Shimizu and Terada2014) where they studied protein–ligand binding for two systems, levansucrase-glucose and LinB-1,2-dichloroethane. A more recent example is the application of Martini 3 FF to study protein–ligand binding in T4 Lysozyme with different small molecules and several pharmacologically relevant targets, such as G protein-coupled receptors (GPCRs), kinases and one example of nuclear receptor (Souza et al., Reference Souza, Thallmair, Conflitti, Ramírez-Palacios, Alessandri, Raniolo, Limongelli and Marrink2020), achieving quantitative agreement with experimental binding affinities. Other examples are present in the literature (Delort et al., Reference Delort, Renault, Charlier, Raussin, Martinez and Floquet2017; Ferré et al., Reference Ferré, Louet, Saurel, Delort, Czaplicki, M’Kadmi, Damian, Renault, Cantel, Gavara, Demange, Marie, Fehrentz, Floquet, Milon and Banères2019; Jiang and Zhang, Reference Jiang and Zhang2019; Dandekar and Mondal, Reference Dandekar and Mondal2020; Negami et al., Reference Negami, Shimizu and Terada2020). Furthermore, the application of Gō models (Poma et al., Reference Poma, Cieplak and Theodorakis2017) in the Martini 3 model leads to an improved description of protein flexibility while preserving computational efficiency. Combined with the new Martini 3 small molecule library (Souza et al., Reference Souza, Limongelli, Wu, Marrink and Monticelli2021b; Alessandri et al., Reference Alessandri, Barnoud, Gertsen, Patmanidis, de Vries, Souza and Marrink2022), CG Martini models now gather the conditions for successful applications in structure-based drug discovery campaigns.

In this perspective, we discuss potential applications of Martini 3 CG simulations to topics relevant for drug discovery and development pipelines, including design of innovative therapies, binding site identification and optimisation of soft delivery systems.

Protein conformation and cryptic pockets

Drug-binding sites are usually pockets or grooves located on the surface of the target protein (Vajda et al., Reference Vajda, Beglov, Wakefield, Egbert and Whitty2018) accessible even in the absence of the drug (Vajda et al., Reference Vajda, Beglov, Wakefield, Egbert and Whitty2018). However, since proteins are dynamic objects, ‘hidden’ binding sites may appear in the presence of an interacting compound (Oleinikovas et al., Reference Oleinikovas, Saladino, Cossins and Gervasio2016; Vajda et al., Reference Vajda, Beglov, Wakefield, Egbert and Whitty2018). These cryptic pockets are often not apparent on the unbound protein surface, only transiently opening up as rare events or shaping themselves in the presence of a ligand (Fig. 1). Cryptic sites can provide unforeseen tractable drug target sites, thus expanding the druggable proteome considerably (Vajda et al., Reference Vajda, Beglov, Wakefield, Egbert and Whitty2018; Hopkins and Groom, Reference Hopkins and Groom2002). On one hand, cryptic pockets offer the prospect to design allosteric drugs (Wenthur et al., Reference Wenthur, Gentry, Mathews and Lindsley2014), a strategy that could be exploited as a therapeutic path towards treating cancer (Zhong et al., Reference Zhong, Li, Xiong, Wang, Wu, Yuan, Yang, Tian, Miao, Wang and Yang2021), diabetes (Wang et al., Reference Wang, Yang, Cheng, Yang, Wang, Dai, Cai, Zhang, Yuliantie, Liu, Jiang, Liu, Wang and Yang2021a), and more recently, SARS-CoV-2 infections (Zimmerman et al., Reference Zimmerman, Porter, Ward, Singh, Vithani, Meller, Mallimadugula, Kuhn, Borowsky, Wiewiora, Hurley, Harbison, Fogarty, Coffland, Fadda, Voelz, Chodera and Bowman2021). On the other hand, cryptic pockets commonly occur at protein–protein interfaces [PPI; (see section ‘Drugs targeting protein–protein interactions’)]. Therefore, the ability to discover and target cryptic pockets would also enable the design of compounds targeting PPIs en route to new therapeutic formulations (Wells and McClendon, Reference Wells and McClendon2007; Shan et al., Reference Shan, Mysore, Leffler, Kim, Sagawa and Shaw2022).

Fig. 1. Schematic representation of a GPCR (PDB IDs 5XEZ & 6LMK) in inactive (left) and active (right) conformations with an allosteric and peptide ligand bound, respectively. Large conformational changes occur upon binding of the peptide ligand and Gs-protein binding intracellularly, which represent possible dynamics that could be observed with Martin combined with Gō-models. The allosteric pocket in the transmembrane domain exemplifies the possibility to use Martini models for identifying transmembrane pockets, allosteric or cryptic, in various complex membrane compositions. Once a ligand is bound, backmapping is a possibility to obtain higher resolution information for further ligand optimisation or design. All figures were rendered using VMD (Humphrey et al., Reference Humphrey, Dalke and Schulten1996).

Several approaches have been proposed for the identification of cryptic sites. While some of them are entirely based on the analysis of protein crystallographic structures (Le Guilloux et al., Reference Le Guilloux, Schmidtke and Tuffery2009), the majority use MD for the identification of cryptic sites (Le Guilloux et al., Reference Le Guilloux, Schmidtke and Tuffery2009; Kokh et al., Reference Kokh, Richter, Henrich, Czodrowski, Rippmann and Wade2013; Laurent et al., Reference Laurent, Chavent, Cragnolini, Dahl, Pasquali, Derreumaux, Sansom and Baaden2015; Cimermancic et al., Reference Cimermancic, Weinkam, Rettenmaier, Bichmann, Keedy, Woldeyes, Schneidman-Duhovny, Demerdash, Mitchell, Wells, Fraser and Sali2016; Kuzmanic et al., Reference Kuzmanic, Bowman, Juarez-Jimenez, Michel and Gervasio2020; Zheng, Reference Zheng2021; Shan et al., Reference Shan, Mysore, Leffler, Kim, Sagawa and Shaw2022).

Cryptic sites are not usually captured in the 180,000+ tridimensional structures obtained by state-of-the-art experimental methods (Bank, 2021); their opening generally occurs on the microsecond-to-millisecond time timescale (Kuzmanic et al., Reference Kuzmanic, Bowman, Juarez-Jimenez, Michel and Gervasio2020). These timescales are only accessible to all-atom (AA) MD simulations relying on specialised hardware, like the Anton3 (Dror et al., Reference Dror, Young, Shaw and Padua2011; Shaw et al., Reference Shaw, Adams, Azaria, Bank, Batson, Bell, Bergdorf, Bhatt, Butts, Correia, Dirks, Dror, Eastwood, Edwards, Even, Feldmann, Fenn, Fenton, Forte, Gagliardo, Gill, Gorlatova, Greskamp, Grossman, Gullingsrud, Harper, Hasenplaugh, Heily, Heshmat, Hunt, Ierardi, Iserovich, Jackson, Johnson, Kirk, Klepeis, Kuskin, Mackenzie, Mader, McGowen, McLaughlin, Moraes, Nasr, Nociolo, O’Donnell, Parker, Peticolas, Pocina, Predescu, Quan, Salmon, Schwink, Shim, Siddique, Spengler, Szalay, Tabladillo, Tartler, Taube, Theobald, Towles, Vick, Wang, Wazlowski, Weingarten, Williams and Yuh2021), or massive distributed computing, as in the Folding@home project (Zimmerman et al., Reference Zimmerman, Porter, Ward, Singh, Vithani, Meller, Mallimadugula, Kuhn, Borowsky, Wiewiora, Hurley, Harbison, Fogarty, Coffland, Fadda, Voelz, Chodera and Bowman2021), but not yet for standard GPU-accelerated hardware (Schlick and Portillo-Ledesma, Reference Schlick and Portillo-Ledesma2021). As a workaround, AA MD-based approaches involve the addition of hydrophilic (e.g. acetic acid, isopropanol) or hydrophobic (e.g. benzene) molecules in the simulation, the so-called mixed-solvent MD (Ghanakota and Carlson, Reference Ghanakota and Carlson2016), the addition of the drug in high concentration, the so-called ‘flooding’ MD approach (Amaro and Li, Reference Amaro and Li2010; Gray et al., Reference Gray, Ma, Wagner and van der Vaart2017), or fragment-based screening (MacKerell et al., Reference MacKerell, Jo, Lakkaraju, Lind and Yu2020). However, also in this case the opening of cryptic pockets may still require several microseconds (Kuzmanic et al., Reference Kuzmanic, Bowman, Juarez-Jimenez, Michel and Gervasio2020). Enhanced sampling approaches have also been used (Bono et al., Reference Bono, De Smet, Herbert, De Bock, Georgiadou, Fons, Tjwa, Alcouffe, Ny, Bianciotto, Jonckx, Murakami, Lanahan, Michielsen, Sibrac, Dol-Gleizes, Mazzone, Zacchigna, Herault, Fischer, Rigon, Ruiz de Almodovar, Claes, Blanc, Poesen, Zhang, Segura, Gueguen, Bordes, Lambrechts, Broussy, van de Wouwer, Michaux, Shimada, Jean, Blacher, Noel, Motte, Rom, Rakic, Katsuma, Schaeffer, Yayon, Van Schepdael, Schwalbe, Gervasio, Carmeliet, Rozensky, Dewerchin, Simons, Christopoulos, Herbert and Carmeliet2013; Herbert et al., Reference Herbert, Schieborr, Saxena, Juraszek, De Smet, Alcouffe, Bianciotto, Saladino, Sibrac, Kudlinzki, Sreeramulu, Brown, Rigon, Herault, Lassalle, Blundell, Rousseau, Gils, Schymkowitz, Tompa, Herbert, Carmeliet, Gervasio, Schwalbe and Bono2013; Oleinikovas et al., Reference Oleinikovas, Saladino, Cossins and Gervasio2016). For the collective variable (CV)-based approaches, the central challenge is choosing a suitable CV (Kuzmanic et al., Reference Kuzmanic, Bowman, Juarez-Jimenez, Michel and Gervasio2020). For the CV-independent methods, running many simulations still entails high computational costs (Earl and Deem, Reference Earl and Deem2005; Kokh et al., Reference Kokh, Czodrowski, Rippmann and Wade2016), which constitutes the main limiting step.

The Martini 3 CG FF, with its increased accuracy and expanded coverage of the chemical space (Souza et al., Reference Souza, Alessandri, Barnoud, Thallmair, Faustino, Grünewald, Patmanidis, Abdizadeh, Bruininks, Wassenaar, Kroon, Melcr, Nieto, Corradi, Khan, Domański, Javanainen, Martinez-Seara, Reuter, Best, Vattulainen, Monticelli, Periole, Tieleman, de Vries and Marrink2021a; Alessandri et al., Reference Alessandri, Barnoud, Gertsen, Patmanidis, de Vries, Souza and Marrink2022), represents a competitive alternative for extracting and targeting druggable structures on such timescales and/or predicting ligand–target interactions (Souza et al., Reference Souza, Thallmair, Conflitti, Ramírez-Palacios, Alessandri, Raniolo, Limongelli and Marrink2020, Reference Souza, Limongelli, Wu, Marrink and Monticelli2021b). So far, one of the limitations of the Martini models is the description of proteins’ conformational flexibility (Poma et al., Reference Poma, Cieplak and Theodorakis2017; Souza et al., Reference Souza, Thallmair, Marrink and Mera-Adasme2019), which is fundamentally linked to biological function (Henzler-Wildman and Kern, Reference Henzler-Wildman and Kern2007; Luo, Reference Luo2012; Veesler and Johnson, Reference Veesler and Johnson2012; Campaner et al., Reference Campaner, Rustighi, Zannini, Cristiani, Piazza, Ciani, Kalid, Golan, Baloglu, Shacham, Valsasina, Cucchi, Pippione, Lolli, Giabbai, Storici, Carloni, Rossetti, Benvenuti, Bello, D’Incalci, Cappuzzello, Rosato and Del Sal2017; Hadden et al., Reference Hadden, Perilla, Schlicksup, Venkatakrishnan, Zlotnick and Schulten2018; Matthes et al., Reference Matthes, Massari, Bochicchio, Schorpp, Schilling, Weber, Offermann, Desantis, Wanker, Carloni, Hadian, Tabarrini, Rossetti and Krauss2018; Maggi et al., Reference Maggi, Carloni and Rossetti2020; Bolnykh et al., Reference Bolnykh, Rossetti, Rothlisberger and Carloni2021; Noreng et al., Reference Noreng, Li and Payandeh2021; Jackson et al., Reference Jackson, Farzan, Chen and Choe2022) and pivotal to design new therapeutics (Hammes, Reference Hammes2002; Campaner et al., Reference Campaner, Rustighi, Zannini, Cristiani, Piazza, Ciani, Kalid, Golan, Baloglu, Shacham, Valsasina, Cucchi, Pippione, Lolli, Giabbai, Storici, Carloni, Rossetti, Benvenuti, Bello, D’Incalci, Cappuzzello, Rosato and Del Sal2017; Sengupta and Udgaonkar, Reference Sengupta and Udgaonkar2019; Schulz-Schaeffer et al., Reference Schulz-Schaeffer, Wemheuer, Wrede and Ennaji2020; Gossen et al., Reference Gossen, Albani, Hanke, Joseph, Bergh, Kuzikov, Costanzi, Manelfi, Storici, Gribbon, Beccari, Talarico, Spyrakis, Lindahl, Zaliani, Carloni, Wade, Musiani, Kokh and Rossetti2021; Xiao et al., Reference Xiao, Bondarenko and Wang2021; Zhao et al., Reference Zhao, Capelli, Carloni, Lüscher, Li and Rossetti2021; Margreiter et al., Reference Margreiter, Witzenberger, Wasser, Davydova, Janowski, Metz, Habib, Sahnoun, Sobisch, Poma, Palomino-Hernandez, Wagner, Carell, Jon Shah, Schulz, Niessing, Voigt and Rossetti2022).

Commonly, Martini-based approaches implement an elastic network (EN), that is addition of a network of harmonic restraints to stabilise protein tertiary structure (Periole et al., Reference Periole, Cavalli, Marrink and Ceruso2009). The restraints are usually added based on a distance criterion, introducing a strong bias towards the starting conformation (Periole et al., Reference Periole, Cavalli, Marrink and Ceruso2009). Different strategies were devised to address this issue (Deplazes et al., Reference Deplazes, Louhivuori, Jayatilaka, Marrink and Corry2012; Lelimousin et al., Reference Lelimousin, Limongelli and Sansom2016; Poma et al., Reference Poma, Cieplak and Theodorakis2017): (i) localised distance-restraints on selected secondary structure elements, often driven by experimental information. For example, this approach was used to study the activation of the epidermal growth factor receptor, coupling Martini 2 CG simulations with enhanced sampling techniques, such as well-tempered metadynamics (Barducci et al., Reference Barducci, Bussi and Parrinello2008) and distance-based restraints on transmembrane helices (Lelimousin et al., Reference Lelimousin, Limongelli and Sansom2016). (ii) implementation of Gō-like models (GōMartini) by establishing a Lennard–Jones (LJ) potential based on the contact map of the native protein structure instead of the harmonic based potential (Poma et al., Reference Poma, Cieplak and Theodorakis2017; Souza et al., Reference Souza, Thallmair, Marrink and Mera-Adasme2019; Mahmood et al., Reference Mahmood, Poma and Okazaki2021). Different contact map definitions for GōMartini were tested on three protein systems (cohesin, titin and ubiquitin) and reproduced protein flexibility as observed in AA simulations. Two different contact map definitions were tested and compared against ENs. The first variant only considers van der Waals spheres (OV) overlaps, while the second builds on top of the OV approach and includes chemical information from the atoms in question. Here a contact between residues requires that the number of attractive contacts between atoms be larger than the number of repulsive ones (Wołek et al., Reference Wołek, Gómez-Sicilia and Cieplak2015; Poma et al., Reference Poma, Cieplak and Theodorakis2017).

Other approaches also exist to characterise conformational transitions between two or more conformational states in the presence of ENs (Kim et al., Reference Kim, Jernigan and Chirikjian2002; Miyashita et al., Reference Miyashita, Onuchic and Wolynes2003; Feng et al., Reference Feng, Yang, Kloczkowski and Jernigan2009; Das et al., Reference Das, Gur, Cheng, Jo, Bahar and Roux2014), like gradually switching between two different types of EN connectivity through a switching parameter or the so-called ‘generalised elastic network’ (Poma et al., Reference Poma, Li and Theodorakis2018) which, within a given cut-off, implements a canonical EN with harmonic potentials and above the chosen cutoff instead implements Gō-like contacts to the system. Recently, adaptive ENs have been developed as well (Kanada et al., Reference Kanada, Terayama, Tokuhisa, Matsumoto and Okuno2022). These strategies represent a potential powerful development strategy for Martini models.

The Martini model can be coupled with strategies to introduce proteins’ dynamics, as discussed above, and combined with enhanced sampling approaches to ideally push the system towards the exploration of ‘rare’ events, like cryptic pocket opening. The use of artificial intelligence algorithms to identify and speed-up the slower modes, as done in Bonati et al. (Reference Bonati, Rizzi and Parrinello2020 and Reference Bonati, Piccini and Parrinello2021) could also be exploited to steer cryptic pockets’ opening. This represents a valid and computationally cheaper solution to identify and target cryptic pockets. Once possible pockets are identified at the CG level, the protein structure could be converted into atomistic resolution (Wassenaar et al., Reference Wassenaar, Pluhackova, Böckmann, Marrink and Tieleman2014; Vickery and Stansfeld, Reference Vickery and Stansfeld2021) for further investigation and ligand design within a virtual screening (VS) workflow.

Protein binding pockets in membrane environments

A fast-growing area for Martini simulations is the analysis of protein–ligand interactions in membrane environments, which is experimentally and computationally rather challenging. Here, the ligand may be an endogenous lipid, that is, natively part of the physiological environment, or an exogenous compound targeting an allosteric pocket of a transmembrane protein or at the protein–lipid interface (Fig. 1). In this context, the structural characterisation of the ligand-binding sites in the transmembrane region and ranking based on binding energetics extracted by CG MD simulations is particularly attractive. This is even more so because standard computational approaches for protein–ligand binding like molecular docking, that do not account for the specificity of the membrane environment, that is, the strong hydrophobic character and the competition with native lipids, are prone to fail. Recently, several CG MD investigations of protein–ligand interactions in the transmembrane region of pharmacologically relevant targets have been reported. In general, the common computational strategy involves: (i) binding-site identification and structural characterisation of the protein–ligand complex using, among other methods, unbiased CG MD simulations and ligand-density maps (Ferraro et al., Reference Ferraro, Masetti, Recanatini, Cavalli and Bottegoni2016; Dämgen and Biggin, Reference Dämgen and Biggin2021); (ii) ranking of binding modes by binding affinity calculations based on equilibrium MD (Souza et al., Reference Souza, Thallmair, Conflitti, Ramírez-Palacios, Alessandri, Raniolo, Limongelli and Marrink2020), potential of mean force (PMF), alchemical transformations, metadynamics (Corey et al., Reference Corey, Vickery, Sansom and Stansfeld2019) or binding saturation curves (Ansell et al., Reference Ansell, Curran, Horrell, Pipatpolkai, Letham, Song, Siebold, Stansfeld, Sansom and Corey2021); and (iii) structural refinement of the protein ligand complex via backmapping to atomistic models (Wassenaar et al., Reference Wassenaar, Pluhackova, Böckmann, Marrink and Tieleman2014). Overall, the main advantage of CG modelling is the ability to converge on sampling the protein–ligand conformational space, currently out-of-reach by typical unbiased atomistic simulations. As a result, within the limits of the accuracy of the model, trends in dissociation constants (Kd) and rates (Koff) can potentially be accessed from unbiased MD (Souza et al., Reference Souza, Thallmair, Conflitti, Ramírez-Palacios, Alessandri, Raniolo, Limongelli and Marrink2020, Reference Souza, Limongelli, Wu, Marrink and Monticelli2021b).

The vast majority of CG MD analyses of protein–ligand interactions in membrane environments involve protein–lipid binding based on Martini 2.2 simulations (De Jong et al., Reference De Jong, Singh, Bennett, Arnarez, Wassenaar, Schäfer, Periole, Tieleman and Marrink2013). The use of the Martini 2.2 FF has allowed not only to discern specific versus nonspecific interactions but also to characterise the energetics involved in the binding reaction. Earlier efforts focused on the prediction of the binding site(s) for cholesterol, which is the most abundant endogenous steroid in mammalian cell membranes and was shown to modulate several membrane proteins including ion channels. Using multi-microsecond CG MD simulations of a homology model of the serotonin transporter embedded in a raft-like membrane, Ferraro et al (Reference Ferraro, Masetti, Recanatini, Cavalli and Bottegoni2016) provided evidence of the existence of specific binding sites for cholesterol, identifying a hotspot that largely overlaps with the cholesterol-binding site illuminated by X-ray crystallography of the closely related dopamine transporter (Ferraro et al., Reference Ferraro, Masetti, Recanatini, Cavalli and Bottegoni2016). By combining CG MD simulations and PMF calculations, Ansell et al (Reference Ansell, Curran, Horrell, Pipatpolkai, Letham, Song, Siebold, Stansfeld, Sansom and Corey2021) characterised the interaction between cholesterol and several membrane proteins including an ATP-dependent pump, a sterol receptor/transporter protein and a member of the TRP ion-channel family. A similar analysis of the chemokine receptor 3, a GPCR responsible for trafficking white blood cells, allowed for the identification of six cholesterol-binding sites, suggesting that recognition of cholesterol at these sites may modulate the affinity for agonists/antagonists allosterically via a rigidification of the protein structure (van Aalst et al., Reference van Aalst, Koneri and Wylie2021). Using CG MD simulations and lipid-density maps, Damgen and Biggin (Reference Dämgen and Biggin2021) explored the affinity of cholesterol and different lipid types for the glycine receptor channel in its active and resting states and found that lipids may act as allosteric modulators because their strength of binding strongly depend(s) on the physiological state of the receptor. In a similar study, protein–lipid interactions on the homologous nicotinic acetylcholine receptor were investigated using a complex quasi-neuronal membrane composed of 36 species of lipids, including cholesterol, in a binding competition assay (Sharp and Brannigan, Reference Sharp and Brannigan2021). Interestingly, the CG MD simulations suggested that cholesterol binds to concave inter-subunit sites and polyunsaturated fatty acids prefer convex sites at the outer transmembrane helix M4, while monounsaturated and saturated lipids are enriched at the protein–lipid interface (Sharp and Brannigan, Reference Sharp and Brannigan2021). Recently, the interaction of the anionic lipids cardiolipins with 42 inner membrane proteins from Escherichia. coli has been investigated by CG MD simulations. Overall, >700 independent cardiolipin binding sites were identified and structurally characterised, thus providing a molecular basis for protein–cardiolipin interactions (Corey et al., Reference Corey, Song, Duncan, Ansell, Sansom and Stansfeld2021). In the context of systematic comparative analyses, the method by Ansell et al. (Reference Ansell, Curran, Horrell, Pipatpolkai, Letham, Song, Siebold, Stansfeld, Sansom and Corey2021) for protein–ligand binding affinities based on binding saturation curves appears particularly appealing as a high-throughput approach for binding-site comparison and ranking.

In addition to protein–lipid interactions, a potential area of development for CG simulations involves the exploration of modulatory ligand binding, such as agonists, antagonists and allosteric modulators, to the transmembrane region of proteins. In this case, and unlike for most lipid molecules, a serious difficulty is introduced by the lack of off-the-shelf CG parameters to model the ligand(s). As a result, examples of studies focusing on the allosteric modulation of transmembrane proteins via protein–ligand interactions are still rare in the literature. One of them focused on the investigation of the binding pathway of two orthosteric agonists of the μ-opioid receptor, that is, fentanyl and morphine (Sutcliffe et al., Reference Sutcliffe, Corey, Charlton, Sessions, Henderson and Kelly2021). Using CG MD simulations and free energy calculations, Sutcliffe et al. (Reference Sutcliffe, Corey, Charlton, Sessions, Henderson and Kelly2021) compared the aqueous and lipophilic binding pathways to the orthosteric site and found that the synthetic opioid fentanyl prefers the lipophilic route, which might explain its lower susceptibility to overdose reversal. Since more and more high-resolution structures of relevant pharmacological targets highlight the existence of multiple allosteric sites in the transmembrane region of these proteins (Cerdan et al., Reference Cerdan, Sisquellas, Pereira, Barreto Gomes, Changeux and Cecchini2020), the development of automatic parameterisation tools to facilitate the setup of CG MD simulations, similar to what is currently available for AA MD, is expected to leverage more exploratory analyses of protein–ligand interactions in the membrane environment and open to high-throughput screening powered by CG MD simulations. Additionally, the new Martini 3 FF offers an extended chemical space (Souza et al., Reference Souza, Thallmair, Conflitti, Ramírez-Palacios, Alessandri, Raniolo, Limongelli and Marrink2020, Reference Souza, Alessandri, Barnoud, Thallmair, Faustino, Grünewald, Patmanidis, Abdizadeh, Bruininks, Wassenaar, Kroon, Melcr, Nieto, Corradi, Khan, Domański, Javanainen, Martinez-Seara, Reuter, Best, Vattulainen, Monticelli, Periole, Tieleman, de Vries and Marrink2021a, Reference Souza, Limongelli, Wu, Marrink and Monticelli2021b), providing an excellent platform for developing automatic parameterisation tools for ligands.

Drugs targeting protein–protein interactions

PPIs have been considered as promising drug targets since the early 2000s, with the hope to overcome the decline in the efficiency of conventional drug development. Three major types of PPI modulators currently described in the literature are small molecules, antibodies and peptides (Mabonga and Kappo, Reference Mabonga and Kappo2019; Lu et al., Reference Lu, Zhou, He, Jiang, Peng, Tong and Shi2020; Martino et al., Reference Martino, Chiarugi, Margheriti and Garau2021). Small molecules typically require a prototypical binding site. The PPI interface is usually flat, shallow and hydrophobic, without an actual pocket where small-molecule ligands may bind (Lu et al., Reference Lu, Zhou, He, Jiang, Peng, Tong and Shi2020). The natural alternative would be to increase the size of the modulator to maximise PPI interface coverage and establish many hydrophobic contacts (Lu et al., Reference Lu, Zhou, He, Jiang, Peng, Tong and Shi2020). However, increasing small-molecule-based PPI modulator size may lead to undesirable pharmacokinetic profiles (An and Fu, Reference An and Fu2018; Lu et al., Reference Lu, Zhou, He, Jiang, Peng, Tong and Shi2020; Martino et al., Reference Martino, Chiarugi, Margheriti and Garau2021). Antibodies present an alternative therapeutic avenue, since these can fully cover the PPI interface due to their size (Bojadzic and Buchwald, Reference Bojadzic and Buchwald2018; Martino et al., Reference Martino, Chiarugi, Margheriti and Garau2021) and there is potential for the general application of antibody-based therapies when combined with novel drug delivery systems (Slastnikova et al., Reference Slastnikova, Ulasov, Rosenkranz and Sobolev2018). Peptides can also be used to modulate PPIs as they bind the PPI interface with high affinity (Cabri et al., Reference Cabri, Cantelmi, Corbisiero, Fantoni, Ferrazzano, Martelli, Mattellone and Tolomelli2021), but they may exhibit short half-lives and toxicity risks (Gupta et al., Reference Gupta, Kapoor, Chaudhary, Gautam, Kumar, Consortium and Raghava2013; Nevola and Giralt, Reference Nevola and Giralt2015; Mabonga and Kappo, Reference Mabonga and Kappo2019). Examples of small-molecule PPI modulators are venetoclax, to treat chronic lymphoblastic leukaemia (Lu et al., Reference Lu, Zhou, He, Jiang, Peng, Tong and Shi2020), and pomalidomide to treat myeloma (Dimopoulos et al., Reference Dimopoulos, Leleu, Palumbo, Moreau, Delforge, Cavo, Ludwig, Morgan, Davies, Sonneveld, Schey, Zweegman, Hansson, Weisel, Mateos, Facon and Miguel2014). ALRN-6924 is an α-helical peptide aimed at leukaemia therapy (Carvajal et al., Reference Carvajal, Neriah, Senecal, Benard, Thiruthuvanathan, Yatsenko, Narayanagari, Wheat, Todorova, Mitchell, Kenworthy, Guerlavais, Annis, Bartholdy, Will, Anampa, Mantzaris, Aivado, Singer, Coleman, Verma and Steidl2018) while Bavencio is an antibody-based drug targeting Merkel cell carcinoma (Boyerinas et al., Reference Boyerinas, Jochems, Fantini, Heery, Gulley, Tsang and Schlom2015).

Recently, a new type of PPI modulator technology gained momentum: the Proteolysis Targeting Chimeras (PROTACs) (Sakamoto et al., Reference Sakamoto, Kim, Kumagai, Mercurio, Crews and Deshaies2001). These bivalent molecules consist of a linker connecting a small molecule binding the target (i.e. ‘warhead’) and a second small molecule binding an E3 ligase (the ‘recruiter’), acting as a PPI enhancer like molecular glues (Wang et al., Reference Wang, Jiang, Feng, Liu and Sun2020; Alabi and Crews, Reference Alabi and Crews2021; Bond and Crews, Reference Bond and Crews2021; Békés et al., Reference Békés, Langley and Crews2022). Simultaneous binding of both proteins by the PROTAC brings them into proximity, provoking target ubiquitination and posterior degradation by proteasome machinery (Wang et al., Reference Wang, Jiang, Feng, Liu and Sun2020; Alabi and Crews, Reference Alabi and Crews2021; Bond and Crews, Reference Bond and Crews2021; Békés et al., Reference Békés, Langley and Crews2022). Compared to small-molecule inhibitors, PROTACs work catalytically, requiring less compound concentration, having fewer off-target effects and exhibiting improved target selectivity (Troup et al., Reference Troup, Fallan and Baud2020; Alabi and Crews, Reference Alabi and Crews2021; Békés et al., Reference Békés, Langley and Crews2022). In the last years, PROTACs attracted the interest of academic and pharmaceutical companies and currently two molecules developed by Arvinas were forwarded to Phase II clinical trials (Petrylak et al., Reference Petrylak, Gao, Vogelzang, Garfield, Taylor, Dougan Moore, Peck and Burris2020; Békés et al., Reference Békés, Langley and Crews2022). Key steps in PROTAC development include the selection of the E3 ligase to pair with the target of interest (Cecchini et al., Reference Cecchini, Pannilunghi, Tardy and Scapozza2021), the accurate prediction of the ternary complex structure (Zaidman et al., Reference Zaidman, Prilusky and London2020) and linker design (Troup et al., Reference Troup, Fallan and Baud2020; Bemis et al., Reference Bemis, La Clair and Burkart2021).

Computational modelling and simulations can help the rational design of PROTACs (Fig. 2). In the absence of ternary complex crystal structures, which must contain the target, the PROTAC and the ligase, one of the first steps of in silico design is sampling the conformational landscape of the complex, which is achievable by protein–protein docking (Hayashi et al., Reference Hayashi, Matsuzaki, Yanagisawa, Ohue and Akiyama2018; Drummond and Williams, Reference Drummond and Williams2019; Drummond et al., Reference Drummond, Henry, Li and Williams2020; Rosell and Fernández-Recio, Reference Rosell and Fernández-Recio2020; Zaidman et al., Reference Zaidman, Prilusky and London2020; Bluntzer et al., Reference Bluntzer, O’Connell, Baker, Michel and Hulme2021; Bai et al., Reference Bai, Miller, Andrianov, Yates, Kirubakaran and Karanicolas2021, Reference Bai, Riching, Makaju, Wu, Acker, Ou, Zhang, Shen, Bulloch, Rui, Gibson, Daniels, Urh, Rock and Humphreys2022; Weng et al., Reference Weng, Li, Kang and Hou2021) and/or MD simulations at various levels of detail (Rakers et al., Reference Rakers, Bermudez, Keller, Mortier and Wolber2015; Yu et al., Reference Yu, Jo, Lakkaraju, Weber and MacKerell2019; Perez et al., Reference Perez, Perez and Perez2021). For very large systems however, AA MD can become prohibitively expensive (Durrant and McCammon, Reference Durrant and McCammon2011; Amaro et al., Reference Amaro, Baudry, Chodera, Demir, McCammon, Miao and Smith2018; Jung et al., Reference Jung, Kobayashi, Kasahara, Tan, Kuroda, Minami, Ishiduki, Nishiki, Inoue, Ishikawa, Feig and Sugita2021). This is particularly true when considering a VS campaign applied to ternary complexes in explicit solvent, due to system size and complexity. As an alternative, docking and MD simulations based on the Martini 2 and 3 CG framework (Roel-Touris et al., Reference Roel-Touris, Don, Honorato, Rodrigues and Bonvin2019; Roel-Touris and Bonvin, Reference Roel-Touris and Bonvin2020; Souza et al., Reference Souza, Limongelli, Wu, Marrink and Monticelli2021b) may be used to facilitate the study of these large macromolecular systems. In a first stage, CG protein–protein docking can be used to capture the most important features of the interaction complex, providing many potential binding modes. It can then be combined with long and affordable CG MD simulations to probe complex stability, which is critical for PPI drug discovery. One example of CG-docking methods is HADDOCK (Roel-Touris et al., Reference Roel-Touris, Don, Honorato, Rodrigues and Bonvin2019; Roel-Touris and Bonvin, Reference Roel-Touris and Bonvin2020). A limitation of some docking approaches is the treatment of proteins as rigid bodies (Vakser, Reference Vakser2020; Harmalkar and Gray, Reference Harmalkar and Gray2021). Recently, docking approaches including protein flexibility have been developed, including ‘divide-and-conquer’ (Karaca and Bonvin, Reference Karaca and Bonvin2011) and normal mode analysis-based strategies (May and Zacharias, Reference May and Zacharias2008; Moal and Bates, Reference Moal and Bates2010; Jiménez-García et al., Reference Jiménez-García, Roel-Touris, Romero-Durana, Vidal, Jiménez-González and Fernández-Recio2018; Diaz et al., Reference Diaz, Frezza and Martin2021). Alternatively, GōMartini simulations (Poma et al., Reference Poma, Cieplak and Theodorakis2017) could be used to cheaply produce protein–protein conformations which, after a back-mapping procedure, could be used in ensemble docking (Amaro et al., Reference Amaro, Baudry, Chodera, Demir, McCammon, Miao and Smith2018).

Fig. 2. Important steps in PROTAC design for drug discovery campaigns. (a) Protein–protein docking either at the atomistic (ribbons) or coarse-grained level (red and cyan spheres). The E3 ligase is represented in red and the target protein in blue. (b) Coarse-graining of a small -molecule using the Martini 3 force field. (c) Dynamical motions of the ligase and the target (blue and red arrows, respectively) are important to query ternary complex stability in the presence of the PROTAC (represented as van der Waals spheres). All figures were rendered using VMD (Humphrey et al., Reference Humphrey, Dalke and Schulten1996). The ternary complex structure is from Nowak et al. (Reference Nowak, DeAngelo, Buckley, He, Donovan, An, Safaee, Jedrychowski, Ponthier, Ishoey, Zhang, Mancias, Gray, Bradner and Fischer2018) with the PDB ID code 6BN7.

Massive protein–protein docking for target identification (Zhang et al., Reference Zhang, Tang, Wang and Lai2014) of other PPI modulators can also greatly benefit from the use of CG approaches. In the case of PROTACS, not only the target but also the choice of E3 Ligase is fundamental for the stability of the ternary complex and cell-specific target degradation (Békés et al., Reference Békés, Langley and Crews2022). Only a limited number of E3 ligases have been explored towards PROTAC development (Burslem and Crews, Reference Burslem and Crews2020; Troup et al., Reference Troup, Fallan and Baud2020; Alabi and Crews, Reference Alabi and Crews2021). Examples are the von Hippel-Landau or Cereblon E3 ligases (He et al., Reference He, Khan, Huo, Lv, Zhang, Liu, Yuan, Hromas, Xu, Zheng and Zhou2020; Bricelj et al., Reference Bricelj, Steinebach, Kuchta, Gütschow and Sosič2021). However, some ligases, for which there is currently no crystal PROTAC ternary complex available, are known to be enriched in specific cell types (Békés et al., Reference Békés, Langley and Crews2022). Combining CG docking between a PROTAC-containing target and several candidate ligases separately with subsequent CG MD simulations could help to identify the most suitable target-ligase pair, enabling cell-type-based therapeutic PROTAC approaches.

Another PROTAC-specific challenge is the design of the linker portion (Alabi and Crews, Reference Alabi and Crews2021) as there exist no common practices or guidelines, and linker size and flexibility affect the degradation efficiency of PROTACs (Cyrus et al., Reference Cyrus, Wehenkel, Choi, Han, Lee, Swanson and Kim2011; Crew et al., Reference Crew, Raina, Dong, Qian, Wang, Vigil, Serebrenik, Hamman, Morgan, Ferraro, Siu, Neklesa, Winkler, Coleman and Crews2018; Troup et al., Reference Troup, Fallan and Baud2020). Optimal linkers should be long and flexible enough to promote a ternary complex orientation that allow ubiquitin transfer to the lysines on the target surface. However, overly flexible linkers may hamper target degradation efficiency (Cecchini et al., Reference Cecchini, Pannilunghi, Tardy and Scapozza2021). CG-based approaches can help linker optimisation. For example, PROTAC CG docking simulations could be used to evaluate the possibility of other PROTAC molecules fitting into the available volume at the binding interface of a ligase/target complex. One route would be by harnessing structural data like the warhead-recruiter distances, extracted from ternary complexes from the Protein Data Bank (Burley et al., Reference Burley, Bhikadiya, Bi, Bittrich, Chen, Crichlow, Christie, Dalenberg, Di Costanzo, Duarte, Dutta, Feng, Ganesan, Goodsell, Ghosh, Green, Guranović, Guzenko, Hudson, Lawson, Liang, Lowe, Namkoong, Peisach, Persikova, Randle, Rose, Rose, Sali, Segura, Sekharan, Shao, Tao, Voigt, Westbrook, Young, Zardecki and Zhuravleva2021) or from protein–protein docking experiments, as constraints. Filtering the predicted complexes using this information in combination with the docking score and other observables would allow retrieval of the best binding poses per system (Zaidman et al., Reference Zaidman, Prilusky and London2020). From the most stable complexes, probed by CG MD, a linker template could then be designed and subsequently used in VS campaigns targeting chemically diverse linker libraries. Further, chemical modifications around the linker template would enable fine tuning of PROTAC properties like solubility, lipophilicity or toxicity effects (Troup et al., Reference Troup, Fallan and Baud2020). Recently, the group of Kihlberg illustrated that PROTACs cell permeability is deeply related to the linkers’ conformational flexibility. Although these compounds do not conform to oral bioavailability defined by the Lipinski rule-of-5 (Lipinski et al., Reference Lipinski, Lombardo, Dominy and Feeney2001), by acting as ‘molecular chameleons’ they are able to fold-in on themselves in aqueous solution and reduce their solvent-accessible polar surface area to increase cell permeability and then unfurl after crossing the membrane (Atilaw et al., Reference Atilaw, Poongavanam, Svensson Nilsson, Nguyen, Giese, Meibom, Erdelyi and Kihlberg2021). Thus, some of the key factors playing a role in PROTAC cell permeability are linker size (Klein et al., Reference Klein, Townsend, Testa, Zengerle, Maniaci, Hughes, Chan, Ciulli and Lokey2020), polarity and rigidity (Atilaw et al., Reference Atilaw, Poongavanam, Svensson Nilsson, Nguyen, Giese, Meibom, Erdelyi and Kihlberg2021), further highlighting the importance of a rational linker design strategy. Similar concerns related to polarity and membrane permeability are also prevalent in PPI-targeting peptide design (Sugita et al., Reference Sugita, Sugiyama, Fujie, Yoshikawa, Yanagisawa, Ohue and Akiyama2021). As such, transfer free energy calculations carried out at the CG level could enable the direct investigation of the ability of different PPI modulators to cross biological membranes in an efficient and affordable manner while still achieving a high degree of accuracy.

Tuning soft nanoparticles with Martini

Drug efficacy correlates with the ability of the drug to reach the target site in sufficient quantities. A high percentage of approved drugs display low aqueous solubility and are fast degraded. To tackle these problems, delivery systems have been developed (Malmsten, Reference Malmsten2006; Wang et al., Reference Wang, Ye, Gao and Ouyang2021b). Different physicochemical properties of the delivery system, such as morphology, composition and stiffness, can contribute to the drug solubility, targeting efficiency and stability (Zhang et al., Reference Zhang, Gao and Bao2015; Yu et al., Reference Yu, Xu, Tian, Su, Zheng, Yang, Wang, Wang, Zhu, Guo, Zhang, Gan, Shi and Gao2018). As drug carrier rigidity affects physiological membrane crossing, developing soft nanoparticle (SN) systems that can easily deform appears attractive.

SNs include carriers consisting of lipids, polymers or surfactants. Lipid-based carriers are generally biocompatible and highly permeable; however, they exhibit low mechanical stability (Sercombe et al., Reference Sercombe, Veerati, Moheimani, Wu, Sood and Hua2015). Polymer-based carriers, on the other hand, have higher mechanical stability but lower biocompatibility and permeability (Jana et al., Reference Jana, Shyam, Singh, Jayaprakash and Dev2021). It is also possible to combine lipids and polymers and to harness the advantages of each component (Reimhult and Virk, Reference Reimhult and Virk2021). Studies have shown that the mechanism of delivery for SNs depends on their morphology and composition, which is in turn correlated with the distribution of the drug within the carrier (El Maghraby et al., Reference El Maghraby, Barry and Williams2008). However, little is known about the morphology and mechanism of delivery for these hybrid systems (Reimhult and Virk, Reference Reimhult and Virk2021).

CG modelling is a valuable tool for investigating the formation of SNs, their morphology, drug distribution within the carrier and the mechanism of delivery, including the interaction with different biological membranes (Yang et al., Reference Yang, Guo, Tian and Chen2021; Parchekani et al., Reference Parchekani, Allahverdi, Taghdir and Naderi-Manesh2022). The first obstacle for using CG models is constructing the system. Fortunately, an increasing number of tools have been developed for building such CG models, examples being TS2CG, Charmm-GUI and Nano Disc builder, allowing the construction of vesicles and other SNs for drug delivery (Qi et al., Reference Qi, Ingólfsson, Cheng, Lee, Marrink and Im2015; Hsu et al., Reference Hsu, Bruininks, Jefferies, Souza, Lee, Patel, Marrink, Qi, Khalid and Im2017; Kjølbye et al., Reference Kjølbye, De Maria, Wassenaar, Abdizadeh, Marrink, Ferkingoff-Borg and Schiøtt2020; Pezeshkian et al., Reference Pezeshkian, König, Wassenaar and Marrink2020), the Polyply package for constructing polymer-based systems (Grünewald et al., Reference Grünewald, Alessandri, Kroon, Monticelli, Souza and Marrink2022) or the Insane.py script for bilayers (Wassenaar et al., Reference Wassenaar, Ingólfsson, Böckmann, Tieleman and Marrink2015). In combination, protocols for simulating soft delivery systems have also started to appear in the literature (Bruininks et al., Reference Bruininks, Souza and Marrink2019). Several CG studies have been performed using the Martini 2 model, investigating the morphology, size and internal organisation of the different components in lipid and polymer-based carriers (Hashemzadeh et al., Reference Hashemzadeh, Javadi and Darvishi2020; Bono et al., Reference Bono, Coloma Smith, Moreschi, Redaelli, Gautieri and Candiani2021; Gao et al., Reference Gao, Nicolas and Ha-Duong2021). Among the first described SNs are liposomes, consisting of a lipid bilayer surrounding a hydrophilic core, capable of trapping both hydrophobic and hydrophilic drugs. Liposomes were the first delivery system to reach clinical application (Doxil) (James et al., Reference James, Coker, Tomlinson, Harris, Gompels, Pinching and Stewart1994) and have been widely used and characterised for many different therapeutics (Allen and Cullis, Reference Allen and Cullis2013). Further development of liposomes resulted in cationic lipids and subsequently cationic polymers for delivery of NAs, which proved invaluable at the outbreak of the COVID-19 pandemic (Polack et al., Reference Polack, Thomas, Kitchin, Absalon, Gurtman, Lockhart, Perez, Pérez Marc, Moreira, Zerbini, Bailey, Swanson, Roychoudhury, Koury, Li, Kalina, Cooper, Frenck, Hammitt, Türeci, Nell, Schaefer, Ünal, Tresnan, Mather, Dormitzer, Şahin, Jansen and Gruber2020; Baden et al., Reference Baden, El Sahly, Essink, Kotloff, Frey, Novak, Diemert, Spector, Rouphael, Creech, McGettigan, Khetan, Segall, Solis, Brosz, Fierro, Schwartz, Neuzil, Corey, Gilbert, Janes, Follmann, Marovich, Mascola, Polakowski, Ledgerwood, Graham, Bennett, Pajon, Knightly, Leav, Deng, Zhou, Han, Ivarsson, Miller and Zaks2021). Cationic lipids or polymers can condense NA efficiently, thanks to the electrostatic interaction with the negatively charged NA to form lipoplexes and polyplexes, respectively (Li and Szoka, Reference Li and Szoka2007; Schlich et al., Reference Schlich, Palomba, Costabile, Mizrahy, Pannuzzo, Peer and Decuzzi2021). Accurate description of the electrostatic interactions is a major challenge in the case of highly charged lipo- or polyplexes. The challenge could be tackled by developing polarisable Martini models, so far only available for water, ions and proteins with the Martini 2 FF (Yesylevskyy et al., Reference Yesylevskyy, Schäfer, Sengupta and Marrink2010; De Jong et al., Reference De Jong, Singh, Bennett, Arnarez, Wassenaar, Schäfer, Periole, Tieleman and Marrink2013; Michalowsky et al., Reference Michalowsky, Schäfer, Holm and Smiatek2017, Reference Michalowsky, Zeman, Holm and Smiatek2018; Sahoo et al., Reference Sahoo, Lee and Matysiak2022).

The main drawback of permanently charged cationic components is their toxicity and rapid elimination from circulation (Li and Szoka, Reference Li and Szoka2007; Schlich et al., Reference Schlich, Palomba, Costabile, Mizrahy, Pannuzzo, Peer and Decuzzi2021). To avoid the toxicity and increase the circulation time and stability, particles can be covered by a PEGylated lipid shield (Li and Szoka, Reference Li and Szoka2007), although PEGylation has shown to diminish particle uptake in target cells (PEG dilemma) (Gjetting et al., Reference Gjetting, Arildsen, Christensen, Poulsen, Roth, Handlos and Poulsen2010). The Martini model has eased the way to study polymer coating in membranes (Grünewald et al., Reference Grünewald, Rossi, de Vries, Marrink and Monticelli2018; Lemaalem et al., Reference Lemaalem, Hadrioui, Derouiche and Ridouane2020) and NPs (Pannuzzo et al., Reference Pannuzzo, Esposito, Wu, Key, Aryal, Celia, di Marzio, Moghimi and Decuzzi2020).

A step further in the optimisation led to ionisable components, resulting in the formulation of lipid nanoparticles (LNPs) (Schlich et al., Reference Schlich, Palomba, Costabile, Mizrahy, Pannuzzo, Peer and Decuzzi2021) and dendrimers (Palmerston Mendes et al., Reference Palmerston Mendes, Pan and Torchilin2017), branched polymers with well-defined molecular weights. The ionisable components are positively charged at low pH to encapsulate NA, and neutral at higher pH, for example, in the blood, thereby avoiding the drawbacks of lipo- and polyplexes. However, it has been shown that only 2–3% of the nucleotide drug load reaches the cytosol using LNPs (Gilleron et al., Reference Gilleron, Querbes, Zeigerer, Borodovsky, Marsico, Schubert, Manygoats, Seifert, Andree, Stöter, Epstein-Barash, Zhang, Koteliansky, Fitzgerald, Fava, Bickle, Kalaidzidis, Akinc, Maier and Zerial2013). Once the LNP or dendrimer is endocytosed, endosomes eventually fuse with lysosomes and their cargo is degraded. For optimised release, the cargo needs to escape the endosome before fusion with the lysosome (Schlich et al., Reference Schlich, Palomba, Costabile, Mizrahy, Pannuzzo, Peer and Decuzzi2021). A general understanding of the delivery mechanism and its pH dependence is lacking. For investigating pH dependent release routes or interactions with NA, constant pH CG approaches are available (Grünewald et al., Reference Grünewald, Souza, Abdizadeh, Barnoud, de Vries and Marrink2020; Aho et al., Reference Aho, Buslaev, Jansen, Bauer, Groenhof and Hess2022). As a proof of concept, collective interactions between titratable sites in a G5 dendrimer poly(propylene imine) were simulated at different pH values, revealing how the particle expands in radius and increases in degree of protonation with decreasing pH, consistent with previous atomistic studies (Grünewald et al., Reference Grünewald, Souza, Abdizadeh, Barnoud, de Vries and Marrink2020).

The delivery depends on the structural properties of the carrier. For LNPs, two different internal organisations have been proposed based on CG modelling with Martini 2 and cryo-transmission electron microscopy (Leung et al., Reference Leung, Hafez, Baoukina, Belliveau, Zhigaltsev, Afshinmanesh, Tieleman, Hansen, Hope and Cullis2012; Kulkarni et al., Reference Kulkarni, Darjuan, Mercer, Chen, van der Meel, Thewalt, Tam and Cullis2018). Understanding the structure–activity relationship is of paramount importance for the rational design of optimised LNPs and for cell-specific targeting. Cell specific LNPs can be built by changing one or more of the lipid components (Liu et al., Reference Liu, Cheng, Wei, Yu, Johnson, Farbiak and Siegwart2021; Żak and Zangi, Reference Żak and Zangi2021), but models of synthetic lipids are not always available. The combination of the Martini 3 FF, with its extended chemical space, and building tools enables the prediction of properties of both empty structures of SNs, such as LNPs, and complexes with cargo of various sizes, from small interfering RNA (siRNAs) to large messenger RNA (mRNA) molecules, enabling studies of the internal organisation and interactions.

Cell specificity and drug efficacy can, in principle, be optimised in terms of interaction and fusion with the endosomal membrane. To this end, Martini 2 models of complex membranes have previously been constructed (Ingólfsson et al., Reference Ingólfsson, Melo, van Eerden, Arnarez, Lopez, Wassenaar, Periole, de Vries, Tieleman and Marrink2014, Reference Ingólfsson, Bhatia, Zeppelin, Bennett, Carpenter, Hsu, Dharuman, Bremer, Schiøtt, Lightstone and Carpenter2020), which demonstrates the possibility of studying the interaction between various SNs formulations (Lee et al., Reference Lee, Cheng, Yu, Liu, Johnson and Siegwart2021) and cell-specific plasma and endosomal membranes. However, this field remains to be explored.

In perspective, there is a general need to implement alternative design strategies to further optimise SNs. Tuning the chemical groups of the lipids or polymer, along with ratios of components, is key for altering the properties of SNs. However, synthesising and testing combinations of lipid or polymer variables are costly and time consuming. In silico screening of promising formulations is a viable alternative to study the role that each component plays in the SN morphology and delivery process (Fig. 3). One drawback when studying lipid-based systems using CG approaches is the loss of resolution compared to AA representations. For instance, changes in tail composition are not always well captured. Nevertheless, in the new version of Martini different tail chemistries can be easier to represent in the future as a result of the use of small and tiny beads.

Fig. 3. CG modelling enables predictions of organisation, size and stability of SNs containing various building blocks and cargo. Moreover, it can be used to study the interaction between various SN formulations and biological barriers, such as plasma and endosomal membranes. All figures were rendered using VMD (Humphrey et al., Reference Humphrey, Dalke and Schulten1996).

Summary and future directions

For in silico studies of large complex systems, aiming at identifying possible druggable sites, predicting and optimising protein ligand binding for drug design or studying drug delivery systems, the Martini model provides an efficient approach relative to AA MD. Due to the timescales reachable by CG Martini simulations, it is possible to probe systems with respect to pockets formed transiently, interesting for drug discovery campaigns. However, further benchmarking is required to assess the accuracy of the FF. Furthermore, maintaining proteins’ tertiary structure in Martini models requires inclusion of EN or GōMartini potentials. Thus, a reasonable definition of the contact map, from which to draw the network of potentials, is critical. Improvements to the definition of contact maps may consider an ensemble of conformations and use knowledge of hydrogen bonds and residue protonation. Additionally, the use of LJ potentials in GōMartini enables the development of multi-basin models, as previously shown for AA MD (Okazaki et al., Reference Okazaki, Koga, Takada, Onuchic and Wolynes2006), combining Gō-models of different protein conformations to promote conformational transitions.

Currently, a major challenge towards the use of the Martini CG model in drug design and drug delivery is the automatic parameterisation of ligands and components of delivery vectors, including AA-to-CG mapping, construction of the bonded parameters and bead-type assignment. To address this issue, tools like Swarm-CG (Empereur-Mot et al., Reference Empereur-Mot, Pesce, Doni, Bochicchio, Capelli, Perego and Pavan2020) or PyCGTOOL (Graham et al., Reference Graham, Essex and Khalid2017) have been developed. However, these approaches focus solely on optimising the bonded parameters. Automated parameterisation workflows for Martini 2 models of small molecules are available (Bereau and Kremer, Reference Bereau and Kremer2015; Potter et al., Reference Potter, Barrett and Miller2021), including mapping, bonded-parameter definition and bead type selection based on optimisation of oil–water partitioning free energies. However, since the covered chemical space is larger in Martini 3, adapting these codes to Martini 3 is not straightforward. Equally important will be the generation of curated and extended libraries of Martini models, such as MAD (the MArtini Database server – https://mad.ibcp.fr), which can be used as reference to access the accuracy of such automatic approaches (Hilpert et al., Reference Hilpert, Beranger, Souza, Vainikka, Nieto, Marrink, Monticelli and Launay2022). However, the current library of Martini 3 small-molecule models (Alessandri et al., Reference Alessandri, Barnoud, Gertsen, Patmanidis, de Vries, Souza and Marrink2022) may already allow initial benchmarks based on fragment-based strategies. Another challenge is the backmapping procedure from CG to AA resolution (Wassenaar et al., Reference Wassenaar, Pluhackova, Böckmann, Marrink and Tieleman2014; Vickery and Stansfeld, Reference Vickery and Stansfeld2021), as protein side-chain directionality is kept, but the binding mode may not be accurate. A standard solution is to perform cycles of energy minimisation and equilibration on the backmapped structure to improve side-chain packing. Another option in this direction would involve the use of machine learning methods to optimise side-chain orientation (Misiura et al., Reference Misiura, Shroff, Thyer and Kolomeisky2022).

Backmapping and small-molecule automatic parameterisation are fundamental goals towards VS of molecules targeting PPI systems, like PROTACS. Additionally, available tools for CG protein–protein docking with the Martini 3 FF could efficiently provide researchers with reasonable starting structures for these large complexes, whose dynamics can be probed by MD. This is the premise of the currently in-development CG version of LightDock (Roel-Touris et al., Reference Roel-Touris, Bonvin and Jiménez-García2020a, Reference Roel-Touris, Jiménez-García and Bonvin2020b), implementing the Martini 3 FF. Coupling these tools with CG docking and MD simulations would allow to derive rules for PROTAC linker design and screening and/or to evaluate ternary complex stability when varying the Ligase protein. Within the field of drug delivery, the Martini 3 model combined with the implementation of tools and protocols available for constructing and simulating soft delivery systems, such as LNPs, will enable in silico screening of various formulations, permitting more efficient optimisation or rational design of delivery methods. However, for NA-containing drug delivery systems, the parameters for RNA/DNA are still under development in Martini 3 and the lack of experimentally resolved structures complicates FF parameter optimisation. While previous Martini 2 NA models were rather rigid (Uusitalo et al., Reference Uusitalo, Ingólfsson, Akhshi, Tieleman and Marrink2015, Reference Uusitalo, Ingólfsson, Marrink and Faustino2017), improving the dynamics of the future NA Martini 3 models is of utmost importance for the simulations of NA delivery systems.

Overcoming these challenges is fundamental for broader applications of the Martini 3 model in biologically relevant systems like SNs, protein–protein interactions, membrane systems and efficient discovery of druggable cryptic pockets, enabling an even larger impact of CG models in fields of drug discovery and delivery. For validation of the CG modelling within drug discovery, one example is the technique of co-crystallisation or soaking macromolecular crystals, essentially replacing solvent with a ligand within the crystal, enabling the comparison to for example, flooding CG-MD simulations for pocket identification (Wienen-Schmidt et al., Reference Wienen-Schmidt, Oebbeke, Ngo, Heine and Klebe2021). Within the drug delivery field, one could imagine correlating predicted structures and organisation of SNs based on CG-MD simulation with fusion and transfection efficacy (Miao et al., Reference Miao, Lin, Huang, Li, Delcassian, Ge, Shi and Anderson2020) measured experimentally, combined with fluorescence studies (Chen et al., Reference Chen, Zhang, Su, Bao, Xie, Zuo, Yang, Wang, Jiang, Lin and Fang2019) enhancing the understanding and development of such delivery methods.

Open peer review

To view the open peer review materials for this article, please visit http://doi.org/10.1017/qrd.2022.16.

Financial support

BJG is employed by Zymvol Biomodeling on a project which received funding from the European Union’s Horizon 2020 research and innovation programme under Marie Skłodowska-Curie grant agreement No. 801342 (Tecniospring INDUSTRY) and the Government of Catalonia’s Agency for Business Competitiveness (ACCIÓ). AB and MC received funding from the French National Research Agency (Grant no. ANR-18-CE11–0015). LM is supported by the French National Institute of Health and Medical Research (INSERM). PCTS, JM, GPP, and LRK are supported by the French National Center for Scientific Research (CNRS). Further funding of LRK, GPP, PCTS and LM came from a research collaboration with PharmCADD. SA and GR acknowledge the grant from the Interdisciplinary Centre for Clinical Research within the faculty of Medicine at the RWTH Aachen University (IZKF TN1-1/IA 532001; TN1–4/IA 532004) and the Deutsche Forschungsgemeinschaft (DFG) via the Research Training Group RTG2416 MultiSenses-MultiScales (368482240/GRK2416). AM and GR acknowledge the Helmholtz European Partnering fundings for the project ‘Innovative high-performance computing approaches for molecular neuromedicine’. GR acknowledges the Federal Ministry of Education and Research (BMBF) and the state of North Rhine-Westphalia as part of the NHR Program, as well as the Joint Lab ‘Supercomputing and Modeling for the Human Brain’ of the Helmholtz Association, Germany and the two European Union’s Horizon 2020 MSCA Program under grant agreement 956314 [ALLODD].

Conflicts of interest

The authors declare no conflicts of interest.

Footnotes

L.R.K. and G.P.P. contributed equally to this work.

References

Aho, N, Buslaev, P, Jansen, A, Bauer, P, Groenhof, G and Hess, B (2022) Scalable Constant pH Molecular Dynamics in GROMACS. Journal of Chemical Theory and Computation 18, 61486160.CrossRefGoogle ScholarPubMed
Alabi, SB and Crews, CM (2021) Major advances in targeted protein degradation: PROTACs, LYTACs, and MADTACs. The Journal of Biological Chemistry 296, 100647.CrossRefGoogle ScholarPubMed
Alessandri, R, Barnoud, J, Gertsen, AS, Patmanidis, I, de Vries, AH, Souza, PCT and Marrink, SJ (2022) Martini 3 coarse-grained force field: Small molecules. Advanced Theory and Simulations 5, 2100391.CrossRefGoogle Scholar
Alessandri, R, Grünewald, F and Marrink, SJ (2021) The martini model in materials science. Advanced Materials 33, e2008635.CrossRefGoogle ScholarPubMed
Alessandri, R, Souza, PCT, Thallmair, S, Melo, MN, De Vries, AH and Marrink, SJ (2019) Pitfalls of the martini model. Journal of Chemical Theory and Computation 15, 54485460.CrossRefGoogle ScholarPubMed
Allen, TM and Cullis, PR (2013) Liposomal drug delivery systems: From concept to clinical applications. Advanced Drug Delivery Reviews 65, 3648.CrossRefGoogle ScholarPubMed
Amaro, RE, Baudry, J, Chodera, J, Demir, Ö, McCammon, JA, Miao, Y and Smith, JC (2018) Ensemble docking in drug discovery. Biophysical Journal 114, 22712278.CrossRefGoogle ScholarPubMed
Amaro, RE and Li, WW (2010) Emerging methods for ensemble-based virtual screening. Current Topics in Medicinal Chemistry 10, 313.CrossRefGoogle ScholarPubMed
An, S and Fu, L (2018) Small-molecule PROTACs: An emerging and promising approach for the development of targeted therapy drugs. eBioMedicine 36, 553562.CrossRefGoogle ScholarPubMed
Ansell, TB, Curran, L, Horrell, MR, Pipatpolkai, T, Letham, SC, Song, W, Siebold, C, Stansfeld, PJ, Sansom, MSP and Corey, RA (2021) Relative affinities of protein-cholesterol interactions from equilibrium molecular dynamics simulations. Journal of Chemical Theory and Computation 17, 65486558.CrossRefGoogle ScholarPubMed
Atilaw, Y, Poongavanam, V, Svensson Nilsson, C, Nguyen, D, Giese, A, Meibom, D, Erdelyi, M and Kihlberg, J (2021) Solution conformations shed light on PROTAC cell permeability. ACS Medicinal Chemistry Letters 12, 107114.CrossRefGoogle ScholarPubMed
Baden, LR, El Sahly, HM, Essink, B, Kotloff, K, Frey, S, Novak, R, Diemert, D, Spector, SA, Rouphael, N, Creech, CB, McGettigan, J, Khetan, S, Segall, N, Solis, J, Brosz, A, Fierro, C, Schwartz, H, Neuzil, K, Corey, L, Gilbert, P, Janes, H, Follmann, D, Marovich, M, Mascola, J, Polakowski, L, Ledgerwood, J, Graham, BS, Bennett, H, Pajon, R, Knightly, C, Leav, B, Deng, W, Zhou, H, Han, S, Ivarsson, M, Miller, J and Zaks, T (2021) Efficacy and safety of the mRNA-1273 SARS-CoV-2 vaccine. The New England Journal of Medicine 384, 403416.CrossRefGoogle ScholarPubMed
Bai, N, Miller, SA, Andrianov, GV, Yates, M, Kirubakaran, P and Karanicolas, J (2021) Rationalizing PROTAC-mediated ternary complex formation using Rosetta. Journal of Chemical Information and Modeling 61, 13681382.CrossRefGoogle ScholarPubMed
Bai, N, Riching, KM, Makaju, A, Wu, H, Acker, TM, Ou, S-C, Zhang, Y, Shen, X, Bulloch, D, Rui, H, Gibson, B, Daniels, DL, Urh, M, Rock, B and Humphreys, SC (2022) Modeling the CRL4A ligase complex to predict target protein ubiquitination induced by cereblon-recruiting PROTACs. The Journal of Biological Chemistry 298, 101653.CrossRefGoogle ScholarPubMed
Protein Data Bank (2021) PDB Statistics: Overall Growth of Released Structures Per Year. Available at www.rcsb.org/stats/growth/growth-released-structures (accessed April 2022).Google Scholar
Barducci, A, Bussi, G and Parrinello, M (2008) Well-tempered metadynamics: A smoothly converging and tunable free-energy method. Physical Review Letters 100, 020603.CrossRefGoogle ScholarPubMed
Békés, M, Langley, DR and Crews, CM (2022) PROTAC targeted protein degraders: The past is prologue. Nature Reviews. Drug Discovery 21, 181200.CrossRefGoogle ScholarPubMed
Bemis, TA, La Clair, JJ and Burkart, MD (2021) Unraveling the role of linker design in proteolysis targeting chimeras. Journal of Medicinal Chemistry 64, 80428052.CrossRefGoogle ScholarPubMed
Bereau, T and Kremer, K (2015) Automated parametrization of the coarse-grained martini force field for small organic molecules. Journal of Chemical Theory and Computation 11, 27832791.CrossRefGoogle ScholarPubMed
Bluntzer, MTJ, O’Connell, J, Baker, TS, Michel, J and Hulme, AN (2021) Designing stapled peptides to inhibit protein-protein interactions: An analysis of successes in a rapidly changing field. Peptide Science 113, e24191.CrossRefGoogle Scholar
Bojadzic, D and Buchwald, P (2018) Toward small-molecule inhibition of protein-protein interactions: General aspects and recent Progress in targeting costimulatory and Coinhibitory (immune checkpoint) interactions. Current Topics in Medicinal Chemistry 18, 674699.CrossRefGoogle ScholarPubMed
Bollini, M, Domaoal, R, Thakur, V, Gallardo-Macias, R, Spasov, K, Anderson, K and Jorgensen, W (2011) Computationally-guided optimization of a docking hit to yield cathecol diethers as potent anti-HIV agents. Journal of Medicinal Chemistry 54, 85828591.CrossRefGoogle Scholar
Bolnykh, V, Rossetti, G, Rothlisberger, U and Carloni, P (2021) Expanding the boundaries of ligand–target modeling by exascale calculations. Computational Molecular Science 11, e1535.CrossRefGoogle Scholar
Bonati, L, Piccini, G and Parrinello, M (2021) Deep learning the slow modes for rare events sampling. Proceedings of the National Academy of Sciences of the United States of America 118, e2113533118.CrossRefGoogle ScholarPubMed
Bonati, L, Rizzi, V and Parrinello, M (2020) Data-driven collective variables for enhanced sampling. Journal of Physical Chemistry Letters 11, 29983004.CrossRefGoogle ScholarPubMed
Bond, MJ and Crews, CM (2021) Proteolysis targeting chimeras (PROTACs) come of age: Entering the third decade of targeted protein degradation. RSC Chemical Biology 2, 725742.CrossRefGoogle ScholarPubMed
Bono, N, Coloma Smith, B, Moreschi, F, Redaelli, A, Gautieri, A and Candiani, G (2021) In silico prediction of the in vitro behavior of polymeric gene delivery vectors. Nanoscale 13, 83338342.CrossRefGoogle ScholarPubMed
Bono, F, De Smet, F, Herbert, C, De Bock, K, Georgiadou, M, Fons, P, Tjwa, M, Alcouffe, C, Ny, A, Bianciotto, M, Jonckx, B, Murakami, M, Lanahan, AA, Michielsen, C, Sibrac, D, Dol-Gleizes, F, Mazzone, M, Zacchigna, S, Herault, JP, Fischer, C, Rigon, P, Ruiz de Almodovar, C, Claes, F, Blanc, I, Poesen, K, Zhang, J, Segura, I, Gueguen, G, Bordes, MF, Lambrechts, D, Broussy, R, van de Wouwer, M, Michaux, C, Shimada, T, Jean, I, Blacher, S, Noel, A, Motte, P, Rom, E, Rakic, JM, Katsuma, S, Schaeffer, P, Yayon, A, Van Schepdael, A, Schwalbe, H, Gervasio, FL, Carmeliet, G, Rozensky, J, Dewerchin, M, Simons, M, Christopoulos, A, Herbert, JM and Carmeliet, P (2013) Inhibition of tumor angiogenesis and growth by a small-molecule multi-FGF receptor blocker with allosteric properties. Cancer Cell 23, 477488.CrossRefGoogle ScholarPubMed
Boyerinas, B, Jochems, C, Fantini, M, Heery, CR, Gulley, JL, Tsang, KY and Schlom, J (2015) Antibody-dependent cellular cytotoxicity activity of a novel anti-PD-L1 antibody Avelumab (MSB0010718C) on human tumor cells. Cancer Immunology Research 3, 11481157.CrossRefGoogle ScholarPubMed
Bricelj, A, Steinebach, C, Kuchta, R, Gütschow, M and Sosič, I (2021) E3 ligase ligands in successful PROTACs: An overview of syntheses and linker attachment points. Frontiers in Chemistry 9, 707317.CrossRefGoogle ScholarPubMed
Brown, DG and Wobst, HJ (2021) A decade of FDA-approved drugs (2010–2019): Trends and future directions. Journal of Medicinal Chemistry 64, 23122338.CrossRefGoogle ScholarPubMed
Bruininks, BMH, Souza, PCT, Ingólfsson, HI and Marrink, SJ (2020) A molecular view on the escape of lipoplexed DNA from the endosome. eLife 9, e52012.CrossRefGoogle ScholarPubMed
Bruininks, BMH, Souza, PCT and Marrink, SJ (2019) A practical view of the martini force field. Methods in Molecular Biology 2022, 105127.CrossRefGoogle ScholarPubMed
Burley, SK, Bhikadiya, C, Bi, C, Bittrich, S, Chen, L, Crichlow, GV, Christie, CH, Dalenberg, K, Di Costanzo, L, Duarte, JM, Dutta, S, Feng, Z, Ganesan, S, Goodsell, DS, Ghosh, S, Green, RK, Guranović, V, Guzenko, D, Hudson, BP, Lawson, CL, Liang, Y, Lowe, R, Namkoong, H, Peisach, E, Persikova, I, Randle, C, Rose, A, Rose, Y, Sali, A, Segura, J, Sekharan, M, Shao, C, Tao, YP, Voigt, M, Westbrook, JD, Young, JY, Zardecki, C and Zhuravleva, M (2021) RCSB protein data Bank: Powerful new tools for exploring 3D structures of biological macromolecules for basic and applied research and education in fundamental biology, biomedicine, biotechnology, bioengineering and energy sciences. Nucleic Acids Research 49, D437D451.CrossRefGoogle ScholarPubMed
Burslem, GM and Crews, CM (2020) Proteolysis-targeting chimeras as therapeutics and tools for biological discovery. Cell 181, 102114.CrossRefGoogle ScholarPubMed
Cabri, W, Cantelmi, P, Corbisiero, D, Fantoni, T, Ferrazzano, L, Martelli, G, Mattellone, A and Tolomelli, A (2021) Therapeutic peptides targeting PPI in clinical development: Overview, mechanism of action and perspectives. Frontiers in Molecular Biosciences 8, 697586.CrossRefGoogle Scholar
Campaner, E, Rustighi, A, Zannini, A, Cristiani, A, Piazza, S, Ciani, Y, Kalid, O, Golan, G, Baloglu, E, Shacham, S, Valsasina, B, Cucchi, U, Pippione, AC, Lolli, ML, Giabbai, B, Storici, P, Carloni, P, Rossetti, G, Benvenuti, F, Bello, E, D’Incalci, M, Cappuzzello, E, Rosato, A and Del Sal, G (2017) A covalent PIN1 inhibitor selectively targets cancer cells by a dual mechanism of action. Nature Communications 8, 15772.CrossRefGoogle ScholarPubMed
Carvajal, LA, Neriah, DB, Senecal, A, Benard, L, Thiruthuvanathan, V, Yatsenko, T, Narayanagari, S-R, Wheat, JC, Todorova, TI, Mitchell, K, Kenworthy, C, Guerlavais, V, Annis, DA, Bartholdy, B, Will, B, Anampa, JD, Mantzaris, I, Aivado, M, Singer, RH, Coleman, RA, Verma, A and Steidl, U (2018) Dual inhibition of MDMX and MDM2 as a therapeutic strategy in leukemia. Science Translational Medicine 10, eaa03003.CrossRefGoogle ScholarPubMed
Casalini, T (2021) Not only in silico drug discovery: Molecular modeling towards in silico drug delivery formulations. Journal of Controlled Release: Official Journal of the Controlled Release Society 332, 390417.CrossRefGoogle ScholarPubMed
Cecchini, C, Pannilunghi, S, Tardy, S and Scapozza, L (2021) From conception to development: Investigating PROTACs features for improved cell permeability and successful protein degradation. Frontiers in Chemistry 9, 672267.CrossRefGoogle ScholarPubMed
Cerdan, AH, Sisquellas, M, Pereira, G, Barreto Gomes, DE, Changeux, J-P and Cecchini, M (2020) The glycine receptor allosteric ligands library (GRALL). Bioinformatics 36, 33793384.CrossRefGoogle Scholar
Chen, X, Zhang, D, Su, N, Bao, B, Xie, X, Zuo, F, Yang, L, Wang, H, Jiang, L, Lin, Q and Fang, M (2019) Visualizing RNA dynamics in live cells with bright and stable fluorescent RNAs. Nature Biotechnology 37, 12871293.CrossRefGoogle Scholar
Cimermancic, P, Weinkam, P, Rettenmaier, TJ, Bichmann, L, Keedy, DA, Woldeyes, RA, Schneidman-Duhovny, D, Demerdash, ON, Mitchell, JC, Wells, JA, Fraser, JS and Sali, A (2016) CryptoSite: Expanding the druggable proteome by characterization and prediction of cryptic binding sites. Journal of Molecular Biology 428, 709719.CrossRefGoogle ScholarPubMed
Corey, RA, Song, W, Duncan, AL, Ansell, TB, Sansom, MSP and Stansfeld, PJ (2021) Identification and assessment of cardiolipin interactions with inner membrane proteins. Science Advances 7(34), eabh2217.CrossRefGoogle ScholarPubMed
Corey, RA, Vickery, ON, Sansom, MSP and Stansfeld, PJ (2019) Insights into membrane protein-lipid interactions from free energy calculations. Journal of Chemical Theory and Computation 15(10), 57275736.CrossRefGoogle ScholarPubMed
Crew, AP, Raina, K, Dong, H, Qian, Y, Wang, J, Vigil, D, Serebrenik, YV, Hamman, BD, Morgan, A, Ferraro, C, Siu, K, Neklesa, TK, Winkler, JD, Coleman, KG and Crews, CM (2018) Identification and characterization of Von Hippel-Lindau-recruiting proteolysis targeting chimeras (PROTACs) of TANK-binding kinase 1. Journal of Medicinal Chemistry 61, 583598.CrossRefGoogle ScholarPubMed
Crivori, P, Cruciani, G, Carrupt, PA and Testa, B (2000) Predicting blood-brain barrier permeation from three-dimensional molecular structure. Journal of Medicinal Chemistry 43, 22042216.CrossRefGoogle ScholarPubMed
Cyrus, K, Wehenkel, M, Choi, E-Y, Han, H-J, Lee, H, Swanson, H and Kim, K-B (2011) Impact of linker length on the activity of PROTACs. Molecular BioSystems 7, 359364.CrossRefGoogle ScholarPubMed
Dämgen, MA and Biggin, PC (2021) State-dependent protein-lipid interactions of a pentameric ligand-gated ion channel in a neuronal membrane. PLoS Computational Biology 17, e1007856.CrossRefGoogle Scholar
Dandekar, BR and Mondal, J (2020) Capturing protein-ligand recognition pathways in coarse-grained simulation. Journal of Physical Chemistry Letters 11, 53025311.CrossRefGoogle ScholarPubMed
Das, A, Gur, M, Cheng, MH, Jo, S, Bahar, I and Roux, B (2014) Exploring the conformational transitions of biomolecular systems using a simple two-state anisotropic network model. PLoS Computational Biology 10, e1003521.CrossRefGoogle ScholarPubMed
De Jong, DH, Singh, G, Bennett, WFD, Arnarez, C, Wassenaar, TA, Schäfer, LV, Periole, X, Tieleman, DP and Marrink, SJ (2013) Improved parameters for the martini coarse-grained protein force field. Journal of Chemical Theory and Computation 9, 687697.CrossRefGoogle ScholarPubMed
Delort, B, Renault, P, Charlier, L, Raussin, F, Martinez, J and Floquet, N (2017) Coarse-grained prediction of peptide binding to G-protein coupled receptors. Journal of Chemical Information and Modeling 57, 562571.CrossRefGoogle ScholarPubMed
Deplazes, E, Louhivuori, M, Jayatilaka, D, Marrink, SJ and Corry, B (2012) Structural investigation of MscL gating using experimental data and coarse-grained MD simulations. PLoS Computational Biology 8, e1002683.CrossRefGoogle ScholarPubMed
Diaz, NC, Frezza, E and Martin, J (2021) Using normal mode analysis on protein structural models. How far can we go on our predictions? Proteins 89, 531543.CrossRefGoogle Scholar
Dimopoulos, MA, Leleu, X, Palumbo, A, Moreau, P, Delforge, M, Cavo, M, Ludwig, H, Morgan, GJ, Davies, FE, Sonneveld, P, Schey, SA, Zweegman, S, Hansson, M, Weisel, K, Mateos, MV, Facon, T and Miguel, JFS (2014) Expert panel consensus statement on the optimal use of pomalidomide in relapsed and refractory multiple myeloma. Leukemia 28, 15731585.CrossRefGoogle Scholar
Dror, RO, Young, C and Shaw, DE (2011) Anton, a special-purpose molecular simulation machine. In Padua, D (ed.) Encyclopedia of Parallel Computing. Boston, MA: Springer US, pp. 6071.Google Scholar
Drummond, ML, Henry, A, Li, H and Williams, CI (2020) Improved accuracy for modeling PROTAC-mediated ternary complex formation and targeted protein degradation via new in silico methodologies. Journal of Chemical Information and Modeling 60, 52345254.CrossRefGoogle ScholarPubMed
Drummond, ML and Williams, CI (2019) In silico modeling of PROTAC-mediated ternary complexes: Validation and application. Journal of Chemical Information and Modeling 59, 16341644.CrossRefGoogle ScholarPubMed
Durrant, JD and McCammon, JA (2011) Molecular dynamics simulations and drug discovery. BMC Biology 9, 71.CrossRefGoogle ScholarPubMed
Earl, DJ and Deem, MW (2005) Parallel tempering: Theory, applications, and new perspectives. Physical Chemistry Chemical Physics: PCCP 7, 39103916.CrossRefGoogle ScholarPubMed
El Maghraby, GM, Barry, BW and Williams, AC (2008) Liposomes and skin: From drug delivery to model membranes. European Journal of Pharmaceutical Sciences: Official Journal of the European Federation for Pharmaceutical Sciences 34, 203222.CrossRefGoogle ScholarPubMed
Empereur-Mot, C, Pesce, L, Doni, G, Bochicchio, D, Capelli, R, Perego, C and Pavan, GM (2020) Swarm-CG: Automatic parametrization of bonded terms in MARTINI-based coarse-grained models of simple to complex molecules via fuzzy self-tuning particle swarm optimization. ACS Omega 5, 3282332843.CrossRefGoogle ScholarPubMed
Esang, M and Gupta, M (2021) Aducanumab as a novel treatment for Alzheimer’s disease: A decade of hope, controversies, and the future. Cureus 13, e17591.Google ScholarPubMed
Falzone, L, Salomone, S and Libra, M (2018) Evolution of cancer pharmacological treatments at the turn of the third millennium. Frontiers in Pharmacology 9, 1300.CrossRefGoogle ScholarPubMed
Feng, Y, Yang, L, Kloczkowski, A and Jernigan, RL (2009) The energy profiles of atomic conformational transition intermediates of adenylate kinase. Proteins 77, 551558.CrossRefGoogle ScholarPubMed
Ferraro, M, Masetti, M, Recanatini, M, Cavalli, A and Bottegoni, G (2016) Mapping cholesterol interaction sites on serotonin transporter through coarse-grained molecular dynamics. PLoS One 11, e0166196.CrossRefGoogle ScholarPubMed
Ferré, G, Louet, M, Saurel, O, Delort, B, Czaplicki, G, M’Kadmi, C, Damian, M, Renault, P, Cantel, S, Gavara, L, Demange, P, Marie, J, Fehrentz, J-A, Floquet, N, Milon, A and Banères, JL (2019) Structure and dynamics of G protein-coupled receptor-bound ghrelin reveal the critical role of the octanoyl chain. Pro1ceedings of the National Academy of Sciences of the United States of America 116, 1752517530.CrossRefGoogle ScholarPubMed
Franco-Ulloa, S, Guarnieri, D, Riccardi, L, Pompa, PP and De Vivo, M (2021) Association mechanism of peptide-coated metal nanoparticles with model membranes: A coarse-grained study. Journal of Chemical Theory and Computation 17, 45124523.CrossRefGoogle ScholarPubMed
Frey, K, Gray, W, Spasov, K, Bollini, M, Gallardo-Macias, R, Jorgensen, W and Anderson, K (2013) Structure-based evaluation of C5 derivatives in the catechol Diether series targeting HIV-1 reverse transcriptase. Chemical Biology & Drug Design 83, 541549.CrossRefGoogle Scholar
Gao, P, Nicolas, J and Ha-Duong, T (2021) Supramolecular organization of polymer prodrug nanoparticles revealed by coarse-grained simulations. Journal of the American Chemical Society 143, 1741217423.CrossRefGoogle ScholarPubMed
Ghanakota, P and Carlson, HA (2016) Moving beyond active-site detection: MixMD applied to allosteric systems. The Journal of Physical Chemistry. B 120, 86858695.CrossRefGoogle ScholarPubMed
Gilleron, J, Querbes, W, Zeigerer, A, Borodovsky, A, Marsico, G, Schubert, U, Manygoats, K, Seifert, S, Andree, C, Stöter, M, Epstein-Barash, H, Zhang, L, Koteliansky, V, Fitzgerald, K, Fava, E, Bickle, M, Kalaidzidis, Y, Akinc, A, Maier, M and Zerial, M (2013) Image-based analysis of lipid nanoparticle–mediated siRNA delivery, intracellular trafficking and endosomal escape. Nature Biotechnology 31, 638646.CrossRefGoogle ScholarPubMed
Gjetting, T, Arildsen, NS, Christensen, CL, Poulsen, TT, Roth, JA, Handlos, VN and Poulsen, HS (2010) In vitro and in vivo effects of polyethylene glycol (PEG)-modified lipid in DOTAP/cholesterol-mediated gene transfection. International Journal of Nanomedicine 5, 371383.Google ScholarPubMed
Gossen, J, Albani, S, Hanke, A, Joseph, BP, Bergh, C, Kuzikov, M, Costanzi, E, Manelfi, C, Storici, P, Gribbon, P, Beccari, AR, Talarico, C, Spyrakis, F, Lindahl, E, Zaliani, A, Carloni, P, Wade, RC, Musiani, F, Kokh, DB and Rossetti, G (2021) A blueprint for high affinity SARS-CoV-2 Mpro inhibitors from activity-based compound library screening guided by analysis of protein dynamics. ACS Pharmacology & Translational Science 4, 10791095.CrossRefGoogle ScholarPubMed
Graham, JA, Essex, JW and Khalid, S (2017) PyCGTOOL: Automated generation of coarse-grained molecular dynamics models from atomistic trajectories. Journal of Chemical Information and Modeling 57, 650656.CrossRefGoogle ScholarPubMed
Gray, GM, Ma, N, Wagner, CE and van der Vaart, A (2017) Molecular dynamics simulations and molecular flooding studies of the retinoid X-receptor ligand binding domain. Journal of Molecular Modeling 23, 98.CrossRefGoogle ScholarPubMed
Grünewald, F, Alessandri, R, Kroon, PC, Monticelli, L, Souza, PCT and Marrink, SJ (2022) Polyply; A python suite for facilitating simulations of macromolecules and nanomaterials. Nature Communications 13, 68.CrossRefGoogle ScholarPubMed
Grünewald, F, Rossi, G, de Vries, AH, Marrink, SJ and Monticelli, L (2018) Transferable MARTINI model of poly(ethylene oxide). The Journal of Physical Chemistry. B 122, 74367449.CrossRefGoogle Scholar
Grünewald, F, Souza, PCT, Abdizadeh, H, Barnoud, J, de Vries, AH and Marrink, SJ (2020) Titratable martini model for constant pH simulations. The Journal of Chemical Physics 153, 024118.CrossRefGoogle ScholarPubMed
Gupta, KM, Das, S and Chow, PS (2021) Molecular dynamics simulations to elucidate translocation and permeation of active from lipid nanoparticle to skin: Complemented by experiments. Nanoscale 13, 1291612928.CrossRefGoogle ScholarPubMed
Gupta, S, Kapoor, P, Chaudhary, K, Gautam, A, Kumar, R, Consortium, OSDD and Raghava, GPS (2013) In silico approach for predicting toxicity of peptides and proteins. PLoS One 8, e73957.CrossRefGoogle ScholarPubMed
Hadden, JA, Perilla, JR, Schlicksup, CJ, Venkatakrishnan, B, Zlotnick, A and Schulten, K (2018) All-atom molecular dynamics of the HBV capsid reveals insights into biological function and cryo-EM resolution limits. eLife 7, e32478.CrossRefGoogle ScholarPubMed
Hammes, GG (2002) Multiple conformational changes in enzyme catalysis. Biochemistry 41, 82218228.CrossRefGoogle ScholarPubMed
Harmalkar, A and Gray, JJ (2021) Advances to tackle backbone flexibility in protein docking. Current Opinion in Structural Biology 67, 178186.CrossRefGoogle ScholarPubMed
Hashemzadeh, H, Javadi, H and Darvishi, MH (2020) Study of structural stability and formation mechanisms in DSPC and DPSM liposomes: A coarse-grained molecular dynamics simulation. Scientific Reports 10, 1837.CrossRefGoogle ScholarPubMed
Hayashi, T, Matsuzaki, Y, Yanagisawa, K, Ohue, M and Akiyama, Y (2018) MEGADOCK-web: An integrated database of high-throughput structure-based protein-protein interaction predictions. BMC Bioinformatics 19, 62.CrossRefGoogle ScholarPubMed
He, Y, Khan, S, Huo, Z, Lv, D, Zhang, X, Liu, X, Yuan, Y, Hromas, R, Xu, M, Zheng, G and Zhou, D (2020) Proteolysis targeting chimeras (PROTACs) are emerging therapeutics for hematologic malignancies. Journal of Hematology & Oncology 13, 103.CrossRefGoogle ScholarPubMed
Henzler-Wildman, K and Kern, D (2007) Dynamic personalities of proteins. Nature 450, 964972.CrossRefGoogle ScholarPubMed
Herbert, C, Schieborr, U, Saxena, K, Juraszek, J, De Smet, F, Alcouffe, C, Bianciotto, M, Saladino, G, Sibrac, D, Kudlinzki, D, Sreeramulu, S, Brown, A, Rigon, P, Herault, J-P, Lassalle, G, Blundell, TL, Rousseau, F, Gils, A, Schymkowitz, J, Tompa, P, Herbert, JM, Carmeliet, P, Gervasio, FL, Schwalbe, H and Bono, F (2013) Molecular mechanism of SSR128129E, an extracellularly acting, small-molecule, allosteric inhibitor of FGF receptor signaling. Cancer Cell 23, 489501.CrossRefGoogle ScholarPubMed
Hilpert, C, Beranger, L, Souza, PCT, Vainikka, PA, Nieto, V, Marrink, SJ, Monticelli, L and Launay, G (2022) Facilitating CG simulations with MAD: The MArtini Database Server. BiorXiv. https://doi.org/10.1101/2022.08.03.502585CrossRefGoogle Scholar
Hopkins, AL and Groom, CR (2002) The druggable genome. Nature Reviews. Drug Discovery 1, 727730.CrossRefGoogle ScholarPubMed
Hsu, P-C, Bruininks, BMH, Jefferies, D, Souza, PCT, Lee, J, Patel, DS, Marrink, SJ, Qi, Y, Khalid, S and Im, W (2017) CHARMM-GUI martini maker for modeling and simulation of complex bacterial membranes with lipopolysaccharides. Journal of Computational Chemistry 38, 23542363.CrossRefGoogle ScholarPubMed
Humphrey, W, Dalke, A and Schulten, K (1996) VMD: Visual molecular dynamics. Journal of Molecular Graphics 14, 3338, 27–28.CrossRefGoogle ScholarPubMed
Ingólfsson, HI, Bhatia, H, Zeppelin, T, Bennett, WFD, Carpenter, KA, Hsu, PC, Dharuman, G, Bremer, PT, Schiøtt, B, Lightstone, FC and Carpenter, TS (2020) Capturing biologically complex tissue-specific membranes at different levels of compositional complexity. The Journal of Physical Chemistry. B 124, 78197829.CrossRefGoogle ScholarPubMed
Ingólfsson, HI, Melo, MN, van Eerden, FJ, Arnarez, C, Lopez, CA, Wassenaar, TA, Periole, X, de Vries, AH, Tieleman, DP and Marrink, SJ (2014) Lipid organization of the plasma membrane. Journal of the American Chemical Society 136, 1455414559.CrossRefGoogle ScholarPubMed
Jackson, CB, Farzan, M, Chen, B and Choe, H (2022) Mechanisms of SARS-CoV-2 entry into cells. Nature Reviews. Molecular Cell Biology 23, 320.CrossRefGoogle ScholarPubMed
James, ND, Coker, RJ, Tomlinson, D, Harris, JR, Gompels, M, Pinching, AJ and Stewart, JS (1994) Liposomal doxorubicin (Doxil): An effective new treatment for Kaposi’s sarcoma in AIDS. Clinical Oncology 6, 294296.CrossRefGoogle Scholar
Jana, P, Shyam, M, Singh, S, Jayaprakash, V and Dev, A (2021) Biodegradable polymers in drug delivery and oral vaccination. European Polymer Journal 142, 110155.CrossRefGoogle Scholar
Jiang, Z and Zhang, H (2019) Molecular mechanism of S1P binding and activation of the S1P receptor. Journal of Chemical Information and Modeling 59, 44024412.CrossRefGoogle Scholar
Jiménez-García, B, Roel-Touris, J, Romero-Durana, M, Vidal, M, Jiménez-González, D and Fernández-Recio, J (2018) LightDock: A new multi-scale approach to protein-protein docking. Bioinformatics 34, 4955.CrossRefGoogle ScholarPubMed
Jorgensen, W (2009) Efficient drug lead discovery and optimization. Accounts of Chemical Research 42, 724733.CrossRefGoogle ScholarPubMed
Jung, J, Kobayashi, C, Kasahara, K, Tan, C, Kuroda, A, Minami, K, Ishiduki, S, Nishiki, T, Inoue, H, Ishikawa, Y, Feig, M and Sugita, Y (2021) New parallel computing algorithm of molecular dynamics for extremely huge scale biological systems. Journal of Computational Chemistry 42, 231241.CrossRefGoogle ScholarPubMed
Kanada, R, Terayama, K, Tokuhisa, A, Matsumoto, S and Okuno, Y (2022) Enhanced conformational sampling with an adaptive coarse-grained elastic network model using short-time all-atom molecular dynamics. Journal of Chemical Theory and Computation 18, 20622074.CrossRefGoogle ScholarPubMed
Kanekal, KH and Bereau, T (2019) Resolution limit of data-driven coarse-grained models spanning chemical space. The Journal of Chemical Physics 151, 164106.CrossRefGoogle ScholarPubMed
Karaca, E and Bonvin, AMJJ (2011) A multidomain flexible docking approach to deal with large conformational changes in the modeling of biomolecular complexes. Structure 19, 555565.CrossRefGoogle ScholarPubMed
Kawamoto, S, Liu, H, Miyazaki, Y, Seo, S, Dixit, M, DeVane, R, MacDermaid, C, Fiorin, G, Klein, ML and Shinoda, W (2022) SPICA force field for proteins and peptides. Journal of Chemical Theory and Computation 18, 32043217.CrossRefGoogle ScholarPubMed
Kim, MK, Jernigan, RL and Chirikjian, GS (2002) Efficient generation of feasible pathways for protein conformational transitions. Biophysical Journal 83, 16201630.CrossRefGoogle ScholarPubMed
Kjølbye, LR, De Maria, L, Wassenaar, TA, Abdizadeh, H, Marrink, SJ, Ferkingoff-Borg, J and Schiøtt, B (2020) A generic protocol for constructing molecular models of nanodiscs in silico. Journal of Chemical Information and Modeling 61, 28692883.CrossRefGoogle Scholar
Klein, VG, Townsend, CE, Testa, A, Zengerle, M, Maniaci, C, Hughes, SJ, Chan, K-H, Ciulli, A and Lokey, RS (2020) Understanding and improving the membrane permeability of VH032-based PROTACs. ACS Medicinal Chemistry Letters 11, 17321738.CrossRefGoogle ScholarPubMed
Kokh, DB, Czodrowski, P, Rippmann, F and Wade, RC (2016) Perturbation approaches for exploring protein binding site flexibility to predict transient binding pockets. Journal of Chemical Theory and Computation 12, 41004113.CrossRefGoogle ScholarPubMed
Kokh, DB, Richter, S, Henrich, S, Czodrowski, P, Rippmann, F and Wade, RC (2013) TRAPP: A tool for analysis of transient binding pockets in proteins. Journal of Chemical Information and Modeling 53, 12351252.CrossRefGoogle ScholarPubMed
Kulkarni, JA, Darjuan, MM, Mercer, JE, Chen, S, van der Meel, R, Thewalt, JL, Tam, YYC and Cullis, PR (2018) On the formation and morphology of lipid nanoparticles containing ionizable cationic lipids and siRNA. ACS Nano 12, 47874795.CrossRefGoogle ScholarPubMed
Kuzmanic, A, Bowman, GR, Juarez-Jimenez, J, Michel, J and Gervasio, FL (2020) Investigating cryptic binding sites by molecular dynamics simulations. Accounts of Chemical Research 53, 654661.CrossRefGoogle ScholarPubMed
Laurent, B, Chavent, M, Cragnolini, T, Dahl, ACE, Pasquali, S, Derreumaux, P, Sansom, MSP and Baaden, M (2015) Epock: Rapid analysis of protein pocket dynamics. Bioinformatics 31, 14781480.CrossRefGoogle ScholarPubMed
Le Guilloux, V, Schmidtke, P and Tuffery, P (2009) Fpocket: An open source platform for ligand pocket detection. BMC Bioinformatics 10, 168.CrossRefGoogle ScholarPubMed
Lee, SM, Cheng, Q, Yu, X, Liu, S, Johnson, LT and Siegwart, DJ (2021) A systematic study of unsaturation in lipid nanoparticles leads to improved mRNA transfection in vivo. Angewandte Chemie 60, 58485853.CrossRefGoogle ScholarPubMed
Lee, S-J, Schlesinger, PH, Wickline, SA, Lanza, GM and Baker, NA (2012) Simulation of fusion-mediated nanoemulsion interactions with model lipid bilayers. Soft Matter 8, 30243035.CrossRefGoogle ScholarPubMed
Lelimousin, M, Limongelli, V and Sansom, MSP (2016) Conformational changes in the epidermal growth factor receptor: Role of the transmembrane domain investigated by coarse-grained MetaDynamics free energy calculations. Journal of the American Chemical Society 138, 1061110622.CrossRefGoogle ScholarPubMed
Lemaalem, M, Hadrioui, N, Derouiche, A and Ridouane, H (2020) Structure and dynamics of liposomes designed for drug delivery: Coarse-grained molecular dynamics simulations to reveal the role of lipopolymer incorporation. RSC Advances 10, 37453755.CrossRefGoogle ScholarPubMed
Leung, AKK, Hafez, IM, Baoukina, S, Belliveau, NM, Zhigaltsev, IV, Afshinmanesh, E, Tieleman, DP, Hansen, CL, Hope, MJ and Cullis, PR (2012) Lipid nanoparticles containing siRNA synthesized by microfluidic mixing exhibit an electron-dense nanostructured Core. The Journal of Physical Chemistry. C, Nanomaterials and Interfaces 116, 1844018450.CrossRefGoogle ScholarPubMed
Li, W and Szoka, FC (2007) Lipid-based nanoparticles for nucleic acid delivery. Pharmaceutical Research 24, 438449.CrossRefGoogle ScholarPubMed
Lin, X, Li, X and Lin, X (2020) A review on applications of computational methods in drug screening and design. Molecules 25, 1375.CrossRefGoogle ScholarPubMed
Lipinski, CA, Lombardo, F, Dominy, BW and Feeney, PJ (2001) Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Advanced Drug Delivery Reviews 46, 326.CrossRefGoogle ScholarPubMed
Liu, S, Cheng, Q, Wei, T, Yu, X, Johnson, LT, Farbiak, L and Siegwart, DJ (2021) Membrane-destabilizing ionizable phospholipids for organ-selective mRNA delivery and CRISPR–Cas gene editing. Nature Materials 20, 701710.CrossRefGoogle ScholarPubMed
López, CA, Rzepiela, AJ, de Vries, AH, Dijkhuizen, L, Hünenberger, PH and Marrink, SJ (2009) Martini coarse-grained force field: Extension to carbohydrates. Journal of Chemical Theory and Computation 5, 31953210.CrossRefGoogle ScholarPubMed
Lu, H, Zhou, Q, He, J, Jiang, Z, Peng, C, Tong, R and Shi, J (2020) Recent advances in the development of protein-protein interactions modulators: Mechanisms and clinical trials. Signal Transduction and Targeted Therapy 5, 213.CrossRefGoogle ScholarPubMed
Luo, M (2012) Influenza virus entry. Advances in Experimental Medicine and Biology 726, 201221.CrossRefGoogle ScholarPubMed
Mabonga, L and Kappo, AP (2019) Protein-protein interaction modulators: Advances, successes and remaining challenges. Biophysical Reviews 11, 559581.CrossRefGoogle ScholarPubMed
Machado, M, Barrera, E, Klein, F, Soñora, M, Silva, S and Pantano, S (2019) The SIRAH2.0 force field: Altius, Fortius, Citius. Journal of Chemical Theory and Computation 15, 27192733.CrossRefGoogle ScholarPubMed
Machado, N, Bruininks, BMH, Singh, P, Santos, L d, Pizzol, CD, Dieamant, GC, Kruger, O, Martin, AA, Marrink, SJ, Souza, PCT and Favero, PP (2022) Complex nanoemulsion for vitamin delivery: Droplet organization and interaction with skin membranes. Nanoscale 14, 506514.CrossRefGoogle ScholarPubMed
MacKerell, AD, Jo, S, Lakkaraju, SK, Lind, C and Yu, W (2020) Identification and characterization of fragment binding sites for allosteric ligand design using the site identification by ligand competitive saturation hotspots approach (SILCS-hotspots). Biochimica et Biophysica Acta, General Subjects 1864, 129519.CrossRefGoogle Scholar
Maggi, L, Carloni, P and Rossetti, G (2020) Modeling the allosteric modulation on a G-protein coupled receptor: The case of M2 muscarinic acetylcholine receptor in complex with LY211960. Scientific Reports 10, 3037.CrossRefGoogle ScholarPubMed
Mahmood, MI, Poma, AB and Okazaki, K-I (2021) Optimizing Gō-MARTINI coarse-grained model for F-BAR protein on lipid membrane. Frontiers in Molecular Biosciences 8, 619381.CrossRefGoogle ScholarPubMed
Malmsten, M (2006) Soft drug delivery systems. Soft Matter 2, 760769.CrossRefGoogle ScholarPubMed
Margreiter, MA, Witzenberger, M, Wasser, Y, Davydova, E, Janowski, R, Metz, J, Habib, P, Sahnoun, SEM, Sobisch, C, Poma, B, Palomino-Hernandez, O, Wagner, M, Carell, T, Jon Shah, N, Schulz, JB, Niessing, D, Voigt, A and Rossetti, G (2022) Small-molecule modulators of TRMT2A decrease PolyQ aggregation and PolyQ-induced cell death. Computational and Structural Biotechnology Journal 20, 443458CrossRefGoogle ScholarPubMed
Marrink, SJ, Corradi, V, Souza, PCT, Ingólfsson, HI, Tieleman, DP and Sansom, MSP (2019) Computational modeling of realistic cell membranes. Chemical Reviews 119, 61846226.CrossRefGoogle ScholarPubMed
Marrink, SJ, Monticelli, L, Melo, MN, Alessandri, R, Tieleman, DP and Souza, PCT (2022) Two decades of martini: Better beads, broader scope. WIREs Computational Molecular Science. 12, e1620.Google Scholar
Marrink, SJ, Risselada, HJ, Yefimov, S, Tieleman, DP and De Vries, AH (2007) The MARTINI force field: Coarse grained model for biomolecular simulations. The Journal of Physical Chemistry. B 111, 78127824.CrossRefGoogle ScholarPubMed
Martino, E, Chiarugi, S, Margheriti, F and Garau, G (2021) Mapping, structure and modulation of PPI. Frontiers in Chemistry 9, 718405.CrossRefGoogle ScholarPubMed
Matthes, F, Massari, S, Bochicchio, A, Schorpp, K, Schilling, J, Weber, S, Offermann, N, Desantis, J, Wanker, E, Carloni, P, Hadian, K, Tabarrini, O, Rossetti, G and Krauss, S (2018) Reducing mutant huntingtin protein expression in living cells by a newly identified RNA CAG binder. ACS Chemical Neuroscience 9, 13991408.CrossRefGoogle ScholarPubMed
May, A and Zacharias, M (2008) Energy minimization in low-frequency normal modes to efficiently allow for global flexibility during systematic protein-protein docking. Proteins 70, 794809.CrossRefGoogle ScholarPubMed
Miao, L, Lin, J, Huang, Y, Li, L, Delcassian, D, Ge, Y, Shi, Y and Anderson, DG (2020) Synergistic lipid compositions for albumin receptor mediated delivery of mRNA to the liver. Nature Communications 11, 113.CrossRefGoogle ScholarPubMed
Michalowsky, J, Schäfer, LV, Holm, C and Smiatek, J (2017) A refined polarizable water model for the coarse-grained MARTINI force field with long-range electrostatic interactions. The Journal of Chemical Physics 146, 054501.CrossRefGoogle ScholarPubMed
Michalowsky, J, Zeman, J, Holm, C and Smiatek, J (2018) A polarizable MARTINI model for monovalent ions in aqueous solution. The Journal of Chemical Physics 149, 163319.CrossRefGoogle ScholarPubMed
Misiura, M, Shroff, R, Thyer, R and Kolomeisky, AB (2022) DLPacker: Deep learning for prediction of amino acid side chain conformations in proteins. Proteins 90, 12781290.CrossRefGoogle ScholarPubMed
Miyashita, O, Onuchic, JN and Wolynes, PG (2003) Nonlinear elasticity, proteinquakes, and the energy landscapes of functional transitions in proteins. Proceedings of the National Academy of Sciences of the United States of America 100, 1257012575.CrossRefGoogle ScholarPubMed
Moal, IH and Bates, PA (2010) SwarmDock and the use of normal modes in protein-protein docking. International Journal of Molecular Sciences 11, 36233648.CrossRefGoogle ScholarPubMed
Mohs, RC and Greig, NH (2017) Drug discovery and development: Role of basic biological research. Alzheimer’s & Dementia: The Journal of the Alzheimer’s Association 3, 651657.Google ScholarPubMed
Montalvo-Acosta, JJ and Cecchini, M (2016) Computational approaches to the chemical equilibrium constant in protein-ligand binding. Molecular Informatics 35, 555567.CrossRefGoogle Scholar
Monticelli, L, Kandasamy, SK, Periole, X, Larson, RG, Tieleman, DP and Marrink, SJ (2008) The MARTINI coarse-grained force field: Extension to proteins. Journal of Chemical Theory and Computation 4, 819834.CrossRefGoogle ScholarPubMed
Negami, T, Shimizu, K and Terada, T (2014) Coarse-grained molecular dynamics simulations of protein-ligand binding. Journal of Computational Chemistry 35, 18351845.CrossRefGoogle ScholarPubMed
Negami, T, Shimizu, K and Terada, T (2020) Coarse-grained molecular dynamics simulation of protein conformational change coupled to ligand binding. Chemical Physics Letters 742, 137144.CrossRefGoogle Scholar
Nevola, L and Giralt, E (2015) Modulating protein-protein interactions: The potential of peptides. Chemical Communications 51, 33023315.CrossRefGoogle ScholarPubMed
Nishiga, M, Wang, DW, Han, Y, Lewis, DB and Wu, JC (2020) COVID-19 and cardiovascular disease: From basic mechanisms to clinical perspectives. Nature Reviews. Cardiology 17, 543558.CrossRefGoogle ScholarPubMed
Noreng, S, Li, T and Payandeh, J (2021) Structural pharmacology of voltage-gated sodium channels. Journal of Molecular Biology 433, 166967.CrossRefGoogle ScholarPubMed
Nowak, RP, DeAngelo, SL, Buckley, D, He, Z, Donovan, KA, An, J, Safaee, N, Jedrychowski, MP, Ponthier, CM, Ishoey, M, Zhang, T, Mancias, JD, Gray, NS, Bradner, JE and Fischer, ES (2018) Plasticity in binding confers selectivity in ligand-induced protein degradation. Nature Chemical Biology 14, 706714.CrossRefGoogle ScholarPubMed
Okazaki, K-I, Koga, N, Takada, S, Onuchic, JN and Wolynes, PG (2006) Multiple-basin energy landscapes for large-amplitude conformational motions of proteins: Structure-based molecular dynamics simulations. Proceedings of the National Academy of Sciences of the United States of America 103, 1184411849.CrossRefGoogle ScholarPubMed
Oleinikovas, V, Saladino, G, Cossins, BP and Gervasio, FL (2016) Understanding cryptic pocket formation in protein targets by enhanced sampling simulations. Journal of the American Chemical Society 138, 1425714263.CrossRefGoogle ScholarPubMed
Palmerston Mendes, L, Pan, J and Torchilin, VP (2017) Dendrimers as nanocarriers for nucleic acid and drug delivery in cancer therapy. Molecules 22(9), 1401.CrossRefGoogle ScholarPubMed
Pannuzzo, M, Esposito, S, Wu, L-P, Key, J, Aryal, S, Celia, C, di Marzio, L, Moghimi, SM and Decuzzi, P (2020) Overcoming nanoparticle-mediated complement activation by surface PEG pairing. Nano Letters 20, 43124321.CrossRefGoogle ScholarPubMed
Parchekani, J, Allahverdi, A, Taghdir, M and Naderi-Manesh, H (2022) Design and simulation of the liposomal model by using a coarse-grained molecular dynamics approach towards drug delivery goals. Scientific Reports 12, 2371.CrossRefGoogle ScholarPubMed
Pereira, G, Szwarc, B, Mondragão, MA, Lima, PA and Pereira, F (2018) A ligand-based approach to the discovery of lead-like potassium channel KV 1.3 inhibitors. ChemistrySelect 3, 13521364.CrossRefGoogle Scholar
Perez, JJ, Perez, RA and Perez, A (2021) Computational modeling as a tool to investigate PPI: From drug design to tissue engineering. Frontiers in Molecular Biosciences 8, 681617.CrossRefGoogle ScholarPubMed
Periole, X, Cavalli, M, Marrink, SJ and Ceruso, MA (2009) Combining an elastic network with a coarse-grained molecular force field: Structure, dynamics, and intermolecular recognition. Journal of Chemical Theory and Computation 5, 25312543.CrossRefGoogle ScholarPubMed
Petrylak, DP, Gao, X, Vogelzang, NJ, Garfield, MH, Taylor, I, Dougan Moore, M, Peck, RA and Burris, HA (2020) First-in-human phase I study of ARV-110, an androgen receptor (AR) PROTAC degrader in patients (pts) with metastatic castrate-resistant prostate cancer (mCRPC) following enzalutamide (ENZ) and/or abiraterone (ABI). Journal of Clinical Orthodontics: JCO 38, 35003500.Google Scholar
Pezeshkian, W, König, M, Wassenaar, TA and Marrink, SJ (2020) Backmapping triangulated surfaces to coarse-grained membrane models. Nature Communications 11, 2296.CrossRefGoogle ScholarPubMed
Polack, FP, Thomas, SJ, Kitchin, N, Absalon, J, Gurtman, A, Lockhart, S, Perez, JL, Pérez Marc, G, Moreira, ED, Zerbini, C, Bailey, R, Swanson, KA, Roychoudhury, S, Koury, K, Li, P, Kalina, WV, Cooper, D, Frenck, RW Jr, Hammitt, LL, Türeci, Ö, Nell, H, Schaefer, A, Ünal, S, Tresnan, DB, Mather, S, Dormitzer, PR, Şahin, U, Jansen, KU and Gruber, WC (2020) Safety and efficacy of the BNT162b2 mRNA Covid-19 vaccine. The New England Journal of Medicine 383, 26032615.CrossRefGoogle ScholarPubMed
Poma, AB, Cieplak, M and Theodorakis, PE (2017) Combining the MARTINI and structure-based coarse-grained approaches for the molecular dynamics studies of conformational transitions in proteins. Journal of Chemical Theory and Computation 13, 13661374.CrossRefGoogle ScholarPubMed
Poma, AB, Li, MS and Theodorakis, PE (2018) Generalization of the elastic network model for the study of large conformational changes in biomolecules. Physical Chemistry Chemical Physics: PCCP 20, 1702017028.CrossRefGoogle Scholar
Potter, TD, Barrett, EL and Miller, MA (2021) Automated coarse-grained mapping algorithm for the martini force field and benchmarks for membrane–water partitioning. Journal of Chemical Theory and Computation 17, 57775791.CrossRefGoogle ScholarPubMed
Qi, Y, Ingólfsson, HI, Cheng, X, Lee, J, Marrink, SJ and Im, W (2015) CHARMM-GUI martini maker for coarse-grained simulations with the martini force field. Journal of Chemical Theory and Computation 11, 44864494.CrossRefGoogle ScholarPubMed
Rakers, C, Bermudez, M, Keller, BG, Mortier, J and Wolber, G (2015) Computational close up on protein-protein interactions: How to unravel the invisible using molecular dynamics simulations? Wiley Interdisciplinary Reviews. Computational Molecular Science 5, 345359.CrossRefGoogle Scholar
Reimhult, E and Virk, MM (2021) Hybrid lipopolymer vesicle drug delivery and release systems. Journal of Biomedical Research 35, 301309.CrossRefGoogle ScholarPubMed
Roel-Touris, J and Bonvin, AMJJ (2020) Coarse-grained (hybrid) integrative modeling of biomolecular interactions. Computational and Structural Biotechnology Journal 18, 11821190.CrossRefGoogle ScholarPubMed
Roel-Touris, J, Bonvin, AMJJ and Jiménez-García, B (2020a) LightDock goes information-driven. Bioinformatics 36, 950952.CrossRefGoogle Scholar
Roel-Touris, J, Don, CG, Honorato, RV, Rodrigues, JPGLM and Bonvin, AMJJ (2019) Less is more: Coarse-grained integrative modeling of large biomolecular assemblies with HADDOCK. Journal of Chemical Theory and Computation 15, 63586367.CrossRefGoogle ScholarPubMed
Roel-Touris, J, Jiménez-García, B and Bonvin, AMJJ (2020b) Integrative modeling of membrane-associated protein assemblies. Nature Communications 11, 6210.CrossRefGoogle Scholar
Rosell, M and Fernández-Recio, J (2020) Docking-based identification of small-molecule binding sites at protein-protein interfaces. Computational and Structural Biotechnology Journal 18, 37503761.CrossRefGoogle ScholarPubMed
Sahoo, A, Lee, P and Matysiak, S (2022) Transferable and polarizable coarse grained model for proteins─ProMPT. Journal of Chemical Theory and Computation 18, 50465055.CrossRefGoogle ScholarPubMed
Sakamoto, KM, Kim, KB, Kumagai, A, Mercurio, F, Crews, CM and Deshaies, RJ (2001) Protacs: Chimeric molecules that target proteins to the Skp1-Cullin-F box complex for ubiquitination and degradation. Proceedings of the National Academy of Sciences of the United States of America 98, 85548559.CrossRefGoogle Scholar
Salassi, S, Caselli, L, Cardellini, J, Lavagna, E, Montis, C, Berti, D and Rossi, G (2021) A martini coarse grained model of citrate-capped gold nanoparticles interacting with lipid bilayers. Journal of Chemical Theory and Computation 17, 65976609.CrossRefGoogle ScholarPubMed
Salassi, S, Simonelli, F, Bartocci, A and Rossi, G (2018) A martini coarse-grained model of the calcein fluorescent dye. Journal of Physics D: Applied Physics 51, 384002.CrossRefGoogle Scholar
Schlander, M, Hernandez-Villafuerte, K, Cheng, C-Y, Mestre-Ferrandiz, J and Baumann, M (2021) How much does it cost to research and develop a new drug? A Systematic Review and Assessment. PharmacoEconomics 39, 12431269.Google ScholarPubMed
Schlich, M, Palomba, R, Costabile, G, Mizrahy, S, Pannuzzo, M, Peer, D and Decuzzi, P (2021) Cytosolic delivery of nucleic acids: The case of ionizable lipid nanoparticles. Bioengineering and Translational Medicine 6, e10213.CrossRefGoogle ScholarPubMed
Schlick, T and Portillo-Ledesma, S (2021) Biomolecular modeling thrives in the age of technology. Nature Computational Science 1, 321331.CrossRefGoogle ScholarPubMed
Schulz-Schaeffer, WJ, Wemheuer, WM and Wrede, A (2020) Chapter 21 - prion diseases: Conformational changes of a protein create an unconventional infectious agent. In Ennaji, MM (ed.), Emerging and Reemerging Viral Pathogens, Academic Press, pp. 479488.CrossRefGoogle Scholar
Sengupta, I and Udgaonkar, J (2019) Monitoring site-specific conformational changes in real-time reveals a misfolding mechanism of the prion protein. eLife 8, e44698.CrossRefGoogle ScholarPubMed
Sercombe, L, Veerati, T, Moheimani, F, Wu, SY, Sood, AK and Hua, S (2015) Advances and challenges of liposome assisted drug delivery. Frontiers in Pharmacology 6, 286.CrossRefGoogle ScholarPubMed
Shan, Y, Mysore, VP, Leffler, AE, Kim, ET, Sagawa, S and Shaw, DE (2022) How does a small molecule bind at a cryptic binding site? PLoS Computational Biology 18, e1009817.CrossRefGoogle Scholar
Sharp, L and Brannigan, G (2021) Spontaneous lipid binding to the nicotinic acetylcholine receptor in a native membrane. The Journal of Chemical Physics 154, 185102.CrossRefGoogle Scholar
Shaw, DE, Adams, PJ, Azaria, A, Bank, JA, Batson, B, Bell, A, Bergdorf, M, Bhatt, J, Butts, JA, Correia, T, Dirks, RM, Dror, RO, Eastwood, MP, Edwards, B, Even, A, Feldmann, P, Fenn, M, Fenton, CH, Forte, A, Gagliardo, J, Gill, G, Gorlatova, M, Greskamp, B, Grossman, JP, Gullingsrud, J, Harper, A, Hasenplaugh, W, Heily, M, Heshmat, BC, Hunt, J, Ierardi, DJ, Iserovich, L, Jackson, BL, Johnson, NP, Kirk, MM, Klepeis, JL, Kuskin, JS, Mackenzie, KM, Mader, RJ, McGowen, R, McLaughlin, A, Moraes, MA, Nasr, MH, Nociolo, LJ, O’Donnell, L, Parker, A, Peticolas, JL, Pocina, G, Predescu, C, Quan, T, Salmon, JK, Schwink, C, Shim, KS, Siddique, N, Spengler, J, Szalay, T, Tabladillo, R, Tartler, R, Taube, AG, Theobald, M, Towles, B, Vick, W, Wang, SC, Wazlowski, M, Weingarten, MJ, Williams, JM and Yuh, KA (2021) Anton 3: Twenty microseconds of molecular dynamics simulation before lunch. In Proceedings of the International Conference for High Performance Computing, Networking, Storage and Analysis. New York, NY: Association for Computing Machinery, pp. 111.Google Scholar
Siebenmorgen, T and Zacharias, M (2020) Computational prediction of protein–protein binding affinities. Computational Molecular Science 10, e1448.CrossRefGoogle Scholar
Slastnikova, TA, Ulasov, AV, Rosenkranz, AA and Sobolev, AS (2018) Targeted intracellular delivery of antibodies: The state of the art. Frontiers in Pharmacology 9, 1208.CrossRefGoogle ScholarPubMed
Śledź, P and Caflisch, A (2018) Protein structure-based drug design: From docking to molecular dynamics. Current Opinion in Structural Biology 48, 93102.CrossRefGoogle ScholarPubMed
Sliwoski, G, Kothiwale, S, Meiler, J and Lowe, EW (2014) Computational methods in drug discovery. Pharmacological Reviews 66, 334395.CrossRefGoogle ScholarPubMed
Souza, PCT, Alessandri, R, Barnoud, J, Thallmair, S, Faustino, I, Grünewald, F, Patmanidis, I, Abdizadeh, H, Bruininks, BMH, Wassenaar, TA, Kroon, PC, Melcr, J, Nieto, V, Corradi, V, Khan, HM, Domański, J, Javanainen, M, Martinez-Seara, H, Reuter, N, Best, RB, Vattulainen, I, Monticelli, L, Periole, X, Tieleman, DP, de Vries, AH and Marrink, SJ (2021a) Martini 3: A general purpose force field for coarse-grained molecular dynamics. Nature Methods 18, 382.CrossRefGoogle Scholar
Souza, PCT, Limongelli, V, Wu, S, Marrink, SJ and Monticelli, L (2021b) Perspectives on high-throughput ligand/protein docking with martini MD simulations. Frontiers in Molecular Biosciences 8, 657222.CrossRefGoogle Scholar
Souza, PCT, Thallmair, S, Conflitti, P, Ramírez-Palacios, C, Alessandri, R, Raniolo, S, Limongelli, V and Marrink, SJ (2020) Protein-ligand binding with the coarse-grained martini model. Nature Communications 11, 3714.CrossRefGoogle ScholarPubMed
Souza, PCT, Thallmair, S, Marrink, SJ and Mera-Adasme, R (2019) An allosteric pathway in copper, zinc superoxide dismutase unravels the molecular mechanism of the G93A amyotrophic lateral sclerosis-linked mutation. Journal of Physical Chemistry Letters 10, 77407744.CrossRefGoogle ScholarPubMed
Sugita, M, Sugiyama, S, Fujie, T, Yoshikawa, Y, Yanagisawa, K, Ohue, M and Akiyama, Y (2021) Large-scale membrane permeability prediction of cyclic peptides crossing a lipid bilayer based on enhanced sampling molecular dynamics simulations. Journal of Chemical Information and Modeling 61, 36813695.CrossRefGoogle Scholar
Sutcliffe, KJ, Corey, RA, Charlton, SJ, Sessions, RB, Henderson, G and Kelly, E (2021) Fentanyl binds to the μ-opioid receptor via the lipid membrane and transmembrane helices. BiorXiv. https://doi.org/10.1101/2021.02.04.429703CrossRefGoogle Scholar
Troup, RI, Fallan, C and Baud, MGJ (2020) Current strategies for the design of PROTAC linkers: A critical review. Exploration of Targeted Anti-tumor Therapy 1, 273312.CrossRefGoogle ScholarPubMed
Uusitalo, JJ, Ingólfsson, HI, Akhshi, P, Tieleman, DP and Marrink, SJ (2015) Martini coarse-grained force field: Extension to DNA. Journal of Chemical Theory and Computation 11, 39323945.CrossRefGoogle Scholar
Uusitalo, JJ, Ingólfsson, HI, Marrink, SJ and Faustino, I (2017) Martini coarse-grained force field: Extension to RNA. Biophysical Journal 113, 246256.CrossRefGoogle Scholar
Vajda, S, Beglov, D, Wakefield, AE, Egbert, M and Whitty, A (2018) Cryptic binding sites on proteins: Definition, detection, and druggability. Current Opinion in Chemical Biology 44, 18.CrossRefGoogle ScholarPubMed
Vakser, IA (2020) Challenges in protein docking. Current Opinion in Structural Biology 64, 160165.CrossRefGoogle ScholarPubMed
van Aalst, E, Koneri, J and Wylie, BJ (2021) In silico identification of cholesterol binding motifs in the chemokine receptor CCR3. Membranes 11, 570.CrossRefGoogle ScholarPubMed
Veesler, D and Johnson, JE (2012) Virus maturation. Annual Review of Biophysics 41, 473496.CrossRefGoogle ScholarPubMed
Vickery, ON and Stansfeld, PJ (2021) CG2AT2: An enhanced fragment-based approach for serial multi-scale molecular dynamics simulations. Journal of Chemical Theory and Computation 17, 64726482.CrossRefGoogle ScholarPubMed
Wang, Y, Jiang, X, Feng, F, Liu, W and Sun, H (2020) Degradation of proteins by PROTACs and other strategies. Acta pharmaceutica Sinica. B 10, 207238.CrossRefGoogle ScholarPubMed
Wang, J, Yang, D, Cheng, X, Yang, L, Wang, Z, Dai, A, Cai, X, Zhang, C, Yuliantie, E, Liu, Q, Jiang, H, Liu, H, Wang, M-W and Yang, H (2021a) Allosteric modulators enhancing GLP-1 binding to GLP-1R via a transmembrane site. ACS Chemical Biology 16, 24442452.CrossRefGoogle Scholar
Wang, W, Ye, Z, Gao, H and Ouyang, D (2021b) Computational pharmaceutics - A new paradigm of drug delivery. Journal of Controlled Release: Official Journal of the Controlled Release Society 338, 119136.CrossRefGoogle Scholar
Wassenaar, TA, Ingólfsson, HI, Böckmann, RA, Tieleman, DP and Marrink, SJ (2015) Computational lipidomics with insane: A versatile tool for generating custom membranes for molecular simulations. Journal of Chemical Theory and Computation 11, 21442155.CrossRefGoogle ScholarPubMed
Wassenaar, TA, Pluhackova, K, Böckmann, RA, Marrink, SJ and Tieleman, DP (2014) Going backward: A flexible geometric approach to reverse transformation from coarse grained to atomistic models. Journal of Chemical Theory and Computation 10, 676690.CrossRefGoogle ScholarPubMed
Weerakoon, D, Carbajo, RJ, De Maria, L, Tyrchan, C and Zhao, H (2022) Impact of PROTAC linker plasticity on the solution conformations and dissociation of the ternary complex. Journal of Chemical Information and Modeling 62, 340349.CrossRefGoogle ScholarPubMed
Wells, JA and McClendon, CL (2007) Reaching for high-hanging fruit in drug discovery at protein-protein interfaces. Nature 450, 10011009.CrossRefGoogle ScholarPubMed
Weng, G, Li, D, Kang, Y and Hou, T (2021) Integrative modeling of PROTAC-mediated ternary complexes. Journal of Medicinal Chemistry 64, 1627116281.CrossRefGoogle ScholarPubMed
Wenthur, CJ, Gentry, PR, Mathews, TP and Lindsley, CW (2014) Drugs for allosteric sites on receptors. Annual Review of Pharmacology and Toxicology 54, 165184.CrossRefGoogle ScholarPubMed
Wienen-Schmidt, B, Oebbeke, M, Ngo, K, Heine, A and Klebe, G (2021) Two methods, one goal: Structural differences between cocrystallization and crystal soaking to discover ligand binding poses. ChemMedChem 16, 292300.CrossRefGoogle ScholarPubMed
Wołek, K, Gómez-Sicilia, À and Cieplak, M (2015) Determination of contact maps in proteins: A combination of structural and chemical approaches. The Journal of Chemical Physics 143, p243105.CrossRefGoogle ScholarPubMed
Xiao, J, Bondarenko, V and Wang, Y (2021) Regulation and drug modulation of a voltage-gated sodium channel: Pivotal role of the S4–S5 linker in activation and slow inactivation. Proceedings of the National Academy of Sciences of the United States of America 118, e2102285118.CrossRefGoogle ScholarPubMed
Yang, Z, Guo, Z, Tian, H and Chen, X (2021) Enhancers in polymeric nonviral gene delivery systems. Viewpoints on Digestive Diseases 2, 20200072.Google Scholar
Yesylevskyy, SO, Schäfer, LV, Sengupta, D and Marrink, SJ (2010) Polarizable water model for the coarse-grained MARTINI force field. PLoS Computational Biology 6, 117.CrossRefGoogle ScholarPubMed
Yu, W, Jo, S, Lakkaraju, SK, Weber, DJ and MacKerell, AD (2019) Exploring protein-protein interactions using the site-identification by ligand competitive saturation methodology. Proteins 87, 289301.CrossRefGoogle ScholarPubMed
Yu, W and MacKerell, AD (2017) Computer-aided drug design methods. Methods in Molecular Biology 1520, 85106.CrossRefGoogle ScholarPubMed
Yu, M, Xu, L, Tian, F, Su, Q, Zheng, N, Yang, Y, Wang, J, Wang, A, Zhu, C, Guo, S, Zhang, X, Gan, Y, Shi, X and Gao, H (2018) Rapid transport of deformation-tuned nanoparticles across biological hydrogels and cellular barriers. Nature Communications 9, 2607.CrossRefGoogle ScholarPubMed
Zaidman, D, Prilusky, J and London, N (2020) PRosettaC: Rosetta based modeling of PROTAC mediated ternary complexes. Journal of Chemical Information and Modeling 60, 48944903.CrossRefGoogle ScholarPubMed
Żak, MM and Zangi, L (2021) Lipid nanoparticles for organ-specific mRNA therapeutic delivery. Pharmaceutics 13, 1675.CrossRefGoogle ScholarPubMed
Zhang, S, Gao, H and Bao, G (2015) Physical principles of nanoparticle cellular endocytosis. ACS Nano 9, 86558671.CrossRefGoogle ScholarPubMed
Zhang, C, Tang, B, Wang, Q and Lai, L (2014) Discovery of binding proteins for a protein target using protein-protein docking-based virtual screening. Proteins 82, 24722482.CrossRefGoogle ScholarPubMed
Zhao, Q, Capelli, R, Carloni, P, Lüscher, B, Li, J and Rossetti, G (2021) Enhanced sampling approach to the induced-fit docking problem in protein-ligand binding: The case of mono-ADP-Ribosylation hydrolase inhibitors. Journal of Chemical Theory and Computation 17, 78997911.CrossRefGoogle Scholar
Zheng, W (2021) Predicting cryptic ligand binding sites based on normal modes guided conformational sampling. Proteins 89, 416426.CrossRefGoogle ScholarPubMed
Zhong, L, Li, Y, Xiong, L, Wang, W, Wu, M, Yuan, T, Yang, W, Tian, C, Miao, Z, Wang, T and Yang, S (2021) Small molecules in targeted cancer therapy: Advances, challenges, and future perspectives. Signal Transduction and Targeted Therapy 6, 201.CrossRefGoogle ScholarPubMed
Zimmerman, MI, Porter, JR, Ward, MD, Singh, S, Vithani, N, Meller, A, Mallimadugula, UL, Kuhn, CE, Borowsky, JH, Wiewiora, RP, Hurley, MFD, Harbison, AM, Fogarty, CA, Coffland, JE, Fadda, E, Voelz, VA, Chodera, JD and Bowman, GR (2021) SARS-CoV-2 simulations go exascale to predict dramatic spike opening and cryptic pockets across the proteome. Nature Chemistry 13, 651659.CrossRefGoogle ScholarPubMed
Figure 0

Fig. 1. Schematic representation of a GPCR (PDB IDs 5XEZ & 6LMK) in inactive (left) and active (right) conformations with an allosteric and peptide ligand bound, respectively. Large conformational changes occur upon binding of the peptide ligand and Gs-protein binding intracellularly, which represent possible dynamics that could be observed with Martin combined with Gō-models. The allosteric pocket in the transmembrane domain exemplifies the possibility to use Martini models for identifying transmembrane pockets, allosteric or cryptic, in various complex membrane compositions. Once a ligand is bound, backmapping is a possibility to obtain higher resolution information for further ligand optimisation or design. All figures were rendered using VMD (Humphrey et al., 1996).

Figure 1

Fig. 2. Important steps in PROTAC design for drug discovery campaigns. (a) Protein–protein docking either at the atomistic (ribbons) or coarse-grained level (red and cyan spheres). The E3 ligase is represented in red and the target protein in blue. (b) Coarse-graining of a small -molecule using the Martini 3 force field. (c) Dynamical motions of the ligase and the target (blue and red arrows, respectively) are important to query ternary complex stability in the presence of the PROTAC (represented as van der Waals spheres). All figures were rendered using VMD (Humphrey et al., 1996). The ternary complex structure is from Nowak et al. (2018) with the PDB ID code 6BN7.

Figure 2

Fig. 3. CG modelling enables predictions of organisation, size and stability of SNs containing various building blocks and cargo. Moreover, it can be used to study the interaction between various SN formulations and biological barriers, such as plasma and endosomal membranes. All figures were rendered using VMD (Humphrey et al., 1996).

Review: Towards design of drugs and delivery systems with the Martini coarse-grained model — R0/PR1

Conflict of interest statement

I have no conflicts of interest to declare.

Comments

Comments to Author: This is an excellently written review of the current state-of-the art methods of using Martini 3 to study drugs and potential delivery systems. I found it to be an informative and interesting read, and have no suggestions at this stage to add to improve quality of the manuscript.

Review: Towards design of drugs and delivery systems with the Martini coarse-grained model — R0/PR2

Conflict of interest statement

Reviewer declares none.

Comments

Comments to Author: Kjølbye and co-authors present a very interesting and instructive review on the potential use of the MARTINI (mostly 3) force field to develop new research strategies for drug design and delivering. It is a well written manuscript with numerous recent and useful references. I would have only few comments to help the readership to contextualize the potential of such developments.

1- Throughout the manuscript (and especially in the introduction), it is not always clear what is done with MARTINI2 or MARTINI3. I understand that the authors want to promote MARTINI3 but this reviewer think that MARTINI2 was already useful to get qualitative results. Thus, it would be nice to precise a little bit more which MARTINI version was used.

2- Beyond MARTINI force field, atomistic simulations are still a gold standard to assess protein-ligand interactions and teams like DE Shaw showed recent successes (https://www.prnewswire.com/news-releases/d-e-shaw-research-licenses-first-in-class-therapeutic-for-immunological-diseases-to-lilly-301566618.html).So, it would be useful to balance a bit more the introduction by presenting recent successes from atomistic simulations.

3- In this review the authors present several examples to develop new drugs but seem to mainly focus on proteins or peptides. Would it be possible to highlight/discuss a bit more the development of this force field for small molecules ?For example, what is the position of the authors for the use of MARTINI force field for fragment based screening (like https://pubmed.ncbi.nlm.nih.gov/31030650/) ?

4- In section A- Protein conformation and cryptic pockets, the authors focused a lot on how MARTINI3 may help to highlight cryptic pockets but then how to use the protein structures ? Would it be used in combination with MARTINI3 small molecules - and in this case how the small molecules can find the pocket ? Would it be then backmapped into atomistic system and used for classical docking ?

5- In section C- Drugs targeting protein-protein interactions, the author mentioned the role of the flexibility for the linkers but did not seem tp present/discuss too much the work done with MARTINI forcefield on Intrinsically Disordered Proteins (IDPs) which may complement the references already presented in this section.

6- Overall, the authors present what it will be possible to do with MARTINI3 in term of modeling but do not present how it can then be compared or validated through experiments. Thus, for a broader audience, this information may help to know how to use MARTINI force field to answer specific experimental questions.

Review: Towards design of drugs and delivery systems with the Martini coarse-grained model — R0/PR3

Conflict of interest statement

Reviewer declares none.

Comments

Comments to Author: The perspective article by Kjølbye et al represents a fresh reading of the recent Martini 3 force field for different applications. The authors address several of the current problem in the methodology and I do find quite good the whole article. However, I suggest to include in the introduction a broad description of other methodologies which are concurrent to Martini 3.

In this way it will also to stress the robustness of Martini 3 as one of the advanced tool in biomolecular modelling. Other CG methodologies have shown relative success in the modelling of nucleic acids (OxDNA/OxRNA), as well as proteins (UNRES) and even implemented for drug discovery (CABS) and so on.

Line 189: Please be more specific about what author means by flavors, I guess it refers to different choices of parameters. Describe the parameters and tuning.

Section A, I believe the reader will be benefited by knowing what changes are present in the new Go-Martini with Martini 3 that is not present in the Poma et al JCTC 2017 original work.

Since the original Go-Martini was also parametrized on the basis of AFM data for nanomechanical studies. I brief description of the current nanomechanics studies could be included.

Authors show the relevance of the contact map in the parametrization of the Go-Martini model. A brief comment on the type of schemes such as atomic overlaps, chemical base or any other one which can assist the construction of a protein model should be mentioned.

In regard of the NA potential in Martini 3, I wonder whether the Uusitalo work using EN in Martini 2 will be revised. EN in protein is fundamental to keep secondary structure, in case of NA, the use of EN represent a very poor description of the stability of the NA (e.g. double helix which is a primary structure in NA). This means in the NA with Martini 3 one will expect to remove EN by improving the energetic description. Can the author comment on limitations of the Martini 2 respect to Martini 3 in regard of NA.

Recommendation: Towards design of drugs and delivery systems with the Martini coarse-grained model — R0/PR4

Comments

Comments to Author: Reviewer #1: This is an excellently written review of the current state-of-the art methods of using Martini 3 to study drugs and potential delivery systems. I found it to be an informative and interesting read, and have no suggestions at this stage to add to improve quality of the manuscript.

Reviewer #2: The perspective article by Kjølbye et al represents a fresh reading of the recent Martini 3 force field for different applications. The authors address several of the current problem in the methodology and I do find quite good the whole article. However, I suggest to include in the introduction a broad description of other methodologies which are concurrent to Martini 3.

In this way it will also to stress the robustness of Martini 3 as one of the advanced tool in biomolecular modelling. Other CG methodologies have shown relative success in the modelling of nucleic acids (OxDNA/OxRNA), as well as proteins (UNRES) and even implemented for drug discovery (CABS) and so on.

Line 189: Please be more specific about what author means by flavors, I guess it refers to different choices of parameters. Describe the parameters and tuning.

Section A, I believe the reader will be benefited by knowing what changes are present in the new Go-Martini with Martini 3 that is not present in the Poma et al JCTC 2017 original work.

Since the original Go-Martini was also parametrized on the basis of AFM data for nanomechanical studies. I brief description of the current nanomechanics studies could be included.

Authors show the relevance of the contact map in the parametrization of the Go-Martini model. A brief comment on the type of schemes such as atomic overlaps, chemical base or any other one which can assist the construction of a protein model should be mentioned.

In regard of the NA potential in Martini 3, I wonder whether the Uusitalo work using EN in Martini 2 will be revised. EN in protein is fundamental to keep secondary structure, in case of NA, the use of EN represent a very poor description of the stability of the NA (e.g. double helix which is a primary structure in NA). This means in the NA with Martini 3 one will expect to remove EN by improving the energetic description. Can the author comment on limitations of the Martini 2 respect to Martini 3 in regard of NA.

Reviewer #3: Kjølbye and co-authors present a very interesting and instructive review on the potential use of the MARTINI (mostly 3) force field to develop new research strategies for drug design and delivering. It is a well written manuscript with numerous recent and useful references. I would have only few comments to help the readership to contextualize the potential of such developments.

1- Throughout the manuscript (and especially in the introduction), it is not always clear what is done with MARTINI2 or MARTINI3. I understand that the authors want to promote MARTINI3 but this reviewer think that MARTINI2 was already useful to get qualitative results. Thus, it would be nice to precise a little bit more which MARTINI version was used.

2- Beyond MARTINI force field, atomistic simulations are still a gold standard to assess protein-ligand interactions and teams like DE Shaw showed recent successes (https://www.prnewswire.com/news-releases/d-e-shaw-research-licenses-first-in-class-therapeutic-for-immunological-diseases-to-lilly-301566618.html).So, it would be useful to balance a bit more the introduction by presenting recent successes from atomistic simulations.

3- In this review the authors present several examples to develop new drugs but seem to mainly focus on proteins or peptides. Would it be possible to highlight/discuss a bit more the development of this force field for small molecules ?For example, what is the position of the authors for the use of MARTINI force field for fragment based screening (like https://pubmed.ncbi.nlm.nih.gov/31030650/) ?

4- In section A- Protein conformation and cryptic pockets, the authors focused a lot on how MARTINI3 may help to highlight cryptic pockets but then how to use the protein structures ? Would it be used in combination with MARTINI3 small molecules - and in this case how the small molecules can find the pocket ? Would it be then backmapped into atomistic system and used for classical docking ?

5- In section C- Drugs targeting protein-protein interactions, the author mentioned the role of the flexibility for the linkers but did not seem tp present/discuss too much the work done with MARTINI forcefield on Intrinsically Disordered Proteins (IDPs) which may complement the references already presented in this section.

6- Overall, the authors present what it will be possible to do with MARTINI3 in term of modeling but do not present how it can then be compared or validated through experiments. Thus, for a broader audience, this information may help to know how to use MARTINI force field to answer specific experimental questions.

Recommendation: Towards design of drugs and delivery systems with the Martini coarse-grained model — R1/PR5

Comments

No accompanying comment.