Next Article in Journal
Synthetic Diphenylacetylene-Based Retinoids Induce DNA Damage in Chinese Hamster Ovary Cells without Altering Viability
Next Article in Special Issue
HCnH Anion Chains with n ≤ 8 Are Nonlinear and Their Permanent Dipole Makes Them Potential Candidates for Astronomical Observation
Previous Article in Journal
Analytical Techniques for Phytocannabinoid Profiling of Cannabis and Cannabis-Based Products—A Comprehensive Review
Previous Article in Special Issue
Structural Properties and Magnetic Ground States of 100 Binary d-Metal Oxides Studied by Hybrid Density Functional Methods
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Martini 3 Model of Cellulose Microfibrils: On the Route to Capture Large Conformational Changes of Polysaccharides

by
Rodrigo A. Moreira
1,
Stefan A. L. Weber
2 and
Adolfo B. Poma
3,*
1
Biosystems and Soft Matter Divison, Institute of Fundamental Technological Research, Polish Academy of Sciences, Pawińskiego 5B, 02-106 Warsaw, Poland
2
Max Planck Institute for Polymer Research, Ackermannweg 10, 55128 Mainz, Germany
3
International Center for Research on Innovative Biobased Materials (ICRI-BioM)—International Research Agenda, Lodz University of Technology, Żeromskiego 116, 90-924 Lodz, Poland
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(3), 976; https://doi.org/10.3390/molecules27030976
Submission received: 22 December 2021 / Revised: 25 January 2022 / Accepted: 28 January 2022 / Published: 1 February 2022
(This article belongs to the Special Issue A Commemorative Special Issue Honoring Professor Donald Truhlar)

Abstract

:
High resolution data from all-atom molecular simulations is used to parameterize a Martini 3 coarse-grained (CG) model of cellulose I allomorphs and cellulose type-II fibrils. In this case, elementary molecules are represented by four effective beads centred in the positions of O2, O3, C6, and O6 atoms in the D-glucose cellulose subunit. Non-bonded interactions between CG beads are tuned according to a low statistical criterion of structural deviation using the Martini 3 type of interactions and are capable of being indistinguishable for all studied cases. To maintain the crystalline structure of each single cellulose chain in the microfibrils, elastic potentials are employed to retain the ribbon-like structure in each chain. We find that our model is capable of describing different fibril-twist angles associated with each type of cellulose fibril in close agreement with atomistic simulation. Furthermore, our CG model poses a very small deviation from the native-like structure, making it appropriate to capture large conformational changes such as those that occur during the self-assembly process. We expect to provide a computational model suitable for several new applications such as cellulose self-assembly in different aqueous solutions and the thermal treatment of fibrils of great importance in bioindustrial applications.

1. Introduction

In nature, biomass is a renewable resource, and today it is considered a key material in the circular economy plan. This natural resource is chemically composed of insoluble carbohydrates (e.g., lignocellulose and cellulose) which under the action of several enzymes can be converted into small monomeric subunits and then easily undergo degradation [1]. The extraction and purification at high quality of polysaccharides for industrial applications (mostly due to its inherent biodegradability) [2] are typically carried out by chemical and physical processes, such as contact with ionic liquids [3] or TEMPO-oxidation [4] and ultrasonic separation [5], respectively. These processes aim to destabilize the polar and electrostatic forces between cellulose chains in fibrils.
Molecular Dynamics (MD) simulation is the computational tool capable of investigating the underlying atomistic mechanisms in macromolecular systems at very short length and time scales (e.g., nm and μs). In this regard, the inner hydrophobic structure of cellulose I allomorphs and type-II fibrils has been characterized by MD methods [6,7,8]. As a result, the hydrogen bond (HB) network necessary to build an elementary microfibril has been quantitatively described in the ground state of the microfibrils in solution [7]. Note that a relatively strong presence of the O3–H⋯O5 intrachain and O6–H⋯O3 interchain HBs are relatively established, and it is believed they are responsible for the stability of the layered structures. This result was validated by X-ray and neutron fiber diffraction [9,10] and demonstrated the absence of O–H⋯O interactions between layers and a larger presence of C–H⋯O HBs between cellulose sheets. Thus, it is expected that some C–H⋯O hydrogen bonds and van der Waals forces may contribute to the stability of cellulose I.
In the age of fast computational processing and machine learning (ML), larger and more complex systems in equilibrium can be modeled by advanced all-atom MD and ML tools, such as viral diseases [11,12], material design [13], etc. However, out of equilibrium processes such as large conformational changes of biomolecules, e.g., the unfolding of a protein, self-assembly of polysaccharides in plant cell walls, enzymatic degradation of plastic, etc., are still beyond the state-of-the-art all-atom MD implementations. For such cases, coarse-grained (CG) methodologies were introduced to reduce the number of degrees-of-freedom, which has a direct impact on the number of simulated particles. In addition, they employ soft potentials which allow the use of larger time steps. In general, they are developed on the basis of rigorous statistical mechanics. Thus, they can deal with larger systems and longer time scales than all-atom MD. Today, one can find several CG models that have been employed to model large fluctuations of cellulose microfibrils. For example, Fan et al. [14] developed a one-site CG model for cellulose I β which described fibrils on a 10–500nm length scale. Srinivas et al. [15] presented a one-site CG model derived by the force-matching approach, which captures the amorphous state of cellulose I β and calculated the free energy of transition between the crystalline to amorphous state. Similarly, our [8] CG model with one-bead per D-glucose centered on C4 atoms was able to model cellulose I allomorphs. In addition, the popular Martini force field [16] has been employed to model crystalline native cellulose (i.e., I β ). Other CG models for sugars [17,18,19] are capable of describing accurately complex sugars (e.g., DNA, RNA, etc.) and capture mechanical properties and dynamics, but they can not reach very large length scales, as required by cellulose fibrils (>40 nm). Some of these models rely on implicit water and thus they can not be used in aqueous conditions, therefore, failing to reproduce the experimental conditions necessary for industrial applications. However, a Martini description could be used in combination with other biomolecules in water such as proteins, lipids, nucleic acids, and other biopolymers. In this regard, a generic framework can be derived in terms of the Martini approach.
Here, we present a Martini 3 model for cellulose fibrils, which employs one set of potential parameters, able to differentiate three structures of cellulose, namely I α , I β , and II. In order to describe the inner structure of the flat ribbon of a cellulose chain, we employed a set of harmonic constraints between nearest neighbours.

2. Materials and Methods

2.1. All-Atom MD Simulation

The initial structures of cellulose I allomorphs (I α and I β ) and cellulose II fibrils were generated by the Cellulose-Builder toolkit [20]. In practice, we prepared systems with 36 cellulose chains, each one composed of 100 D-glucose units. The final structures were visualized by VMD [21], and data were postprocessed using its internal protocols. All-atom MD simulations were performed using NAMD [22] version 2.14. The fibrils were modeled using the CHARMM36 force field [23], where the D-glucose denoted as the “BLGC” residue in the force field was used to describe each residue in a given cellulose chain. A triclinic box was used to represent the simulation box, and periodic boundary conditions were implemented in all directions. Solvation of the simulation box by TIP3P water molecules [24] included a buffer distance of 15 Å from the fibril. A total of 71,286, 67,403, and 59,383 water molecules were necessary for each fibrils system, namely I α , I β , and type-II. The system had an average density of 0.1053 atoms.
The equilibration protocol was: (i) 10,000 of energy minimization via conjugate gradient protocol for the solvent, while the solute remained restrained; (ii) unrestrained MD simulation using a time step of 2 fs, and a reference temperature and pressure of 300 K and 1 atm, respectively. For the second step the Langevin thermostat and Piston, which are implemented in NAMD were employed. Long-range electrostatic interactions were computed using the PME [25] methodology; (iii) finally, a production run of 100 ns for each system was started after the target temperature and pressure were achieved in the NVT and NPT ensembles. Cross sections of the fibril structures are represented in Figure 1.

2.2. Coarse-Grained Model: Martini 3

In this part, we employed a versatile CG force field denoted as Martini 3 [27]. This new methodology has been successfully validated in several applications of large conformational changes in complex systems [28,29,30]. Hence, our aim was to build a robust CG model that not only reproduced the basic structural parameters of cellulose I α , I β , and type-II fibrils but also a CG model, which can be transferable between different fibrils. In order to achieve this goal, we searched in the parameter space of all CG Martini bead types. Since each cellulose fibril has different interchain long-range interaction networks (i.e., different packing structures due to HBs) we had to select an elementary fragment of the cellulose chain that represents all the structures. This study used the whole cellulose chain as the basic building block (BB) for all different fibrils.
For this purpose, we modeled each cellulose fibril as an aggregate of the same BB. Each BB consisted of several planes, with each one defined by 4 CG beads denoted as CG1, CG2, CG3, and CG4 (see Figure 2). The orientation between planes changed along the fibril axis as a consequence of our parameterization. Validation of this minimal model was supported by all-atom MD data. There are several ways to parameterize a plane, and here, we used an elementary parametrization known as the simplicial complex homeomorphic (SCH) to a plane [31]. This mathematical construction is defined by triangular patches that optimally cover an entire surface, in our case the cellulose chains. CG simulations showed an average density of 0.0105 atoms/Å 3 , with an average of 51,291 atoms per box.

3. Results

3.1. Mapping CG Structure by All-Atom MD

Figure 3 shows atomistic distributions of distances between chosen CG sites after mapping via the SCH protocol. For all cellulose fibrils, we observe almost a well-defined bell-shape distribution with an exception for O2-O6 and O3-O6, where a bimodal distribution was sampled. In order to maintain a certain degree of parameter transferability between CG models of cellulose fibrils, we considered a distance value between two CG beads consistent with the average of the means of distributions. For those two cases where distributions diverged, we tuned the distance value to reproduce structural consistency in all CG models. The selected parameters are shown in Table 1.

3.2. Parameterization of the CG Model for Several Cellulose Fibrils

CG systems were simulated using GROMACS 2020.2 [32]. A similar equilibration procedure as described in the all-atom MD section was considered. In particular, long-range interactions were computed using Martini 3 standard methodology [27], namely, the reaction-field method [33] using a cutoff of 1.1 nm. Temperature and pressure control were achieved using the V-rescale thermostat [34] and Parrinello-Rahman barostat [35], respectively. Temperature and pressure were set to 300 K and 1 bar for all CG simulations. Equilibration in the NVT and NPT ensembles lasted 0.1 ns and 1 ns, respectively. The positions of two CG beads, namely CG1 and CG2, were constrained and fully relaxed during 1 ns by the end of the equilibration.
The time step for production MD runs was 0.02 ps and, CG-MD trajectories were set to 50 ns for each of the 18 different Martini 3 interaction types (i.e., P1, P6, SP1, SP6, TP1, TP6, N1, N6, SN1, SN6, TN1, TN6, C1, C6, SC1, SC6, TC1, and TC6). In the Martini 3 library, each interaction type is defined by at least three parameters, i.e., σ i j and ϵ i j correspond to the Lennard-Jones (12-6) potential and q for particle charges [16]. In our case, we decided to parametrize the CG3 with a TC1 Martini 3 bead type as it has the smallest interaction with CG water, generally given by W bead type, and other CG beads. According to the all-atom MD distribution of distances between CG centers (see Figure 3), we defined the same elastic constant equal to 30,000 kJ/mol for well-behaved bell-shape cases. Only those flexible bond distances, “O2-O6” and “O3-O6”, required special attention. In this case, we tuned the elastic strength to a lower range of values using 25 kJ/mol, 250 kJ/mol, and 2500 kJ/mol. A cumulative time of about 3 μs was necessary to determine the best model amongst all parameters.
Figure 4 shows the results of the CG parametrization for several bead types in Martini 3 and the elastic bond strength for O2-O6 and O3-O6. The final set of parameters satisfy the criterion of a small deviation of the RMSD. This quantity was averaged over all the CG trajectory, using as a reference the last frame of AA-MD simulation. As a result we selected the bead type SP6 and 2500 kJ/mol for the elastic constant as they follow our structural criterion, and the RMSD remained smaller for all fibrils simultaneously (see Table 2).

3.3. Structural Validation of the Martini 3 Model for Cellulose Fibrils

Table 3 shows the comparison of distance in CG and MD simulations. The Martini 3 model with parameters SP6 bead type and 2500 kJ/mol for O2-O6 and O3-O6 intramolecular distances maintains accurately the internal structure of the cellulose chain in each cellulose fibril type, namely I α , I β , and type-II. Hence, it highlights a transferable model between those three systems.
Our CG model preserves the internal structure of the D-glucose molecules and shows thermal stability in a range of temperature from 300 K–400 K (see Supplementary Material). These results indicate the possibility to study new applications in Martini 3 such as the structural relaxation of fibrils under different temperatures.
We also carried out a RMSD cross-correlation analysis. We computed an RMSD for each pair of frames from AA and CG trajectories and, subsequently, the RMSD average and respective standard deviations (SD) for each pair of trajectories. This combinatorial problem required the computation of 663,522 values of RMSD from a pair of structures. Note that this procedure captures the relative structural changes from two structures at different time points. Table 4 shows the RMSD cross-correlation for AA and CG trajectories. This result validates our Martini 3 model of several cellulose fibrils, as it shows the correct structural description for each crystalline cellulose fibril due to a small deviation in RMSD in the range of 5 Å. In fact it shows a well-equilibrated CG structure respect to its analogous AA system. Furthermore, the larger deviation with respect to other crystals (RMSD > 15 Å) shows that a given CG fibril model retains a structural distance from other cellulose crystals, as expected. Thus no transition or relaxation from one fibril to to another is observed in our simulations.
Figure 5 shows the analysis of one feature in cellulose fibrils: the twist angle. In our case, we define it by the dihedral angle formed by O2 and C6 centers from two subsequent D-glucose molecules. As a result, we observe qualitative agreement between AA-MD and CG-MD representations for cellulose I allomorphs (I α and I β ). However, our CG model is indeed more accurate for the case of cellulose type-II.
A detailed analysis of the distribution of twist angles shows consistency between AA and CG for type-II cellulose with a mean value of the twist angle about −0.42 degrees. In the case of cellulose I allomorphs, the dihedral angle in CG simulations reported about half of the value in AA-MD (see Table 5).

4. Conclusions

In this study, we employed a mathematical construction called geometrical simplicial complex for the determination of an optimal CG mapping scheme. In this context, this approach offers the advantage that a CG model does not require angular (three or four centers) potentials to be represented. However, a clear limitation such as the case of heterogeneity in complex polysaccharides may require a generalization of our approach.
We have tested our approach in two cellulose I allomorphs (i.e., I α and I β ), which are almost indistinguishable at large length scales. Although the local structure was mostly dominated by HBs, we managed to account for a versatile Martini 3 CG model capable of reproducing a single cellulose chain structure using only three uncharged Martini bead types. Furthermore, we also modelled the structure of cellulose type-II, which is substantially different from cellulose I allomorphs. Furthermore, our model captured the atomistic twist angle inherited by the fibril structure in solution for all three structures. Future experiments can address the relevance of the twist as a function of the temperature.
Our model opens the possibility to study large conformational changes of polysaccharides, as it is fully compatible with our earliest implementation of the Gō-Martini approach for the modelling of large conformational changes of proteins [36]. In this regard, a future improvement of this model for larger conformational changes inside a single CG cellulose chain can be achieved by switching some of the interactions between nearest neighbours by a consistent set of Gō-like potentials following standard rules for contact map creation [7,37]. Similar ideas, such as the dual-basin Gō-like model to capture transitions between different fibrils, can be suitable to speed up transitions [38,39].

Supplementary Materials

The following are available online, Figure S1: Histogram of intra and inter residues angles from CG dynamics, Figure S2: Annealing results for I α , I β , type-II cellulose fibrils.

Author Contributions

Conceptualization, R.A.M., S.A.L.W. and A.B.P.; methodology, R.A.M.; formal analysis, R.A.M.; writing—original draft preparation, A.B.P.; writing—review and editing, R.A.M. and S.A.L.W.; supervision, A.B.P.; funding acquisition, A.B.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Science Center, Poland, under grant 2017/26/D/NZ1/ 00466 and the grant MAB PLUS/2019/11 from the Foundation for Polish Science.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. Source code is available in https://github.com/rams-research/cgcellulosefibril, accessed on 27 January 2022.

Acknowledgments

A.P. acknowledges financial support from the National Science Center, Poland, under grant 2017/26/D/NZ1/00466, the grant MAB PLUS/2019/11 from the Foundation for Polish Science, and also computational resources were supported by the PL-GRID infrastructure. Thanks to Joseph L. Baker for commenting and proofreading the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gefen, G.; Anbar, M.; Morag, E.; Lamed, R.; Bayer, E.A. Enhanced cellulose degradation by targeted integration of a cohesin-fused β-glucosidase into the Clostridium thermocellum cellulosome. Proc. Natl. Acad. Sci. USA 2012, 109, 10298–10303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Haider, T.P.; Völker, C.; Kramm, J.; Landfester, K.; Wurm, F.R. Plastics of the future? The impact of biodegradable polymers on the environment and on society. Angew. Chem. Int. Ed. 2019, 58, 50–62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Zhu, S.; Wu, Y.; Chen, Q.; Yu, Z.; Wang, C.; Jin, S.; Ding, Y.; Wu, G. Dissolution of cellulose with ionic liquids and its application: A mini-review. Green Chem. 2006, 8, 325–327. [Google Scholar] [CrossRef]
  4. Cao, X.; Ding, B.; Yu, J.; Al-Deyab, S.S. Cellulose nanowhiskers extracted from TEMPO-oxidized jute fibers. Carbohydr. Polym. 2012, 90, 1075–1080. [Google Scholar] [CrossRef] [PubMed]
  5. Chen, P.; Yu, H.; Liu, Y.; Chen, W.; Wang, X.; Ouyang, M. Concentration effects on the isolation and dynamic rheological behavior of cellulose nanofibers via ultrasonic processing. Cellulose 2013, 20, 149–157. [Google Scholar] [CrossRef]
  6. Maurer, R.J.; Sax, A.F.; Ribitsch, V. Molecular simulation of surface reorganization and wetting in crystalline cellulose I and II. Cellulose 2013, 20, 25–42. [Google Scholar] [CrossRef] [Green Version]
  7. Poma, A.B.; Chwastyk, M.; Cieplak, M. Polysaccharide–protein complexes in a coarse-grained model. J. Phys. Chem. B 2015, 119, 12028–12041. [Google Scholar] [CrossRef]
  8. Poma, A.B.; Chwastyk, M.; Cieplak, M. Coarse-grained model of the native cellulose Iα and the transformation pathways to the Iβ allomorph. Cellulose 2016, 23, 1573–1591. [Google Scholar] [CrossRef] [Green Version]
  9. Nishiyama, Y.; Langan, P.; Chanzy, H. Crystal structure and hydrogen-bonding system in cellulose Iβ from synchrotron X-ray and neutron fiber diffraction. J. Am. Chem. Soc. 2002, 124, 9074–9082. [Google Scholar] [CrossRef]
  10. Nishiyama, Y.; Sugiyama, J.; Chanzy, H.; Langan, P. Crystal structure and hydrogen bonding system in cellulose Iα from synchrotron X-ray and neutron fiber diffraction. J. Am. Chem. Soc. 2003, 125, 14300–14306. [Google Scholar] [CrossRef]
  11. Koehler, M.; Ray, A.; Moreira, R.A.; Juniku, B.; Poma, A.B.; Alsteens, D. Molecular insights into receptor binding energetics and neutralization of SARS-CoV-2 variants. Nat. Commun. 2021, 12, 6977. [Google Scholar] [CrossRef] [PubMed]
  12. Martinez, M.; Cooper, C.D.; Poma, A.B.; Vargas Guzman, H. A Free Energies of the Disassembly of Viral Capsids from a Multiscale Molecular Simulation Approach. J. Chem. Inf. Model 2020, 60, 974–981. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Gkeka, P.; Stoltz, G.; Barati Farimani, A.; Belkacemi, Z.; Ceriotti, M.; Chodera, J.D.; Dinner, A.R.; Ferguson, A.L.; Maillet, J.B.; Minoux, H.; et al. Machine learning force fields and coarse-grained variables in molecular dynamics: Application to materials and biological systems. J. Chem. Theory Comput. 2020, 16, 4757–4775. [Google Scholar] [CrossRef]
  14. Fan, B.; Maranas, J.K. Coarse-grained simulation of cellulose Iβ with application to long fibrils. Cellulose 2015, 22, 31–44. [Google Scholar] [CrossRef]
  15. Srinivas, G.; Cheng, X.; Smith, J.C. Coarse-grain model for natural cellulose fibrils in explicit water. J. Phys. Chem. B 2014, 118, 3026–3034. [Google Scholar] [CrossRef]
  16. Marrink, S.J.; Risselada, H.J.; Yefimov, S.; Tieleman, D.P.; De Vries, A.H. The MARTINI force field: Coarse grained model for biomolecular simulations. J. Phys. Chem. B 2007, 111, 7812–7824. [Google Scholar] [CrossRef] [Green Version]
  17. Liwo, A.; Czaplewski, C.; Sieradzan, A.K.; Lipska, A.G.; Samsonov, S.A.; Murarka, R.K. Theory and practice of coarse-grained molecular dynamics of biologically important systems. Biomolecules 2021, 11, 1347. [Google Scholar] [CrossRef]
  18. Ouldridge, T.E.; Louis, A.A.; Doye, J.P. DNA nanotweezers studied with a coarse-grained model of DNA. Phys. Rev. Lett. 2010, 104, 178101. [Google Scholar] [CrossRef]
  19. Gopal, S.M.; Mukherjee, S.; Cheng, Y.M.; Feig, M. PRIMO/PRIMONA: A coarse-grained model for proteins and nucleic acids that preserves near-atomistic accuracy. Proteins Struct. Funct. Bioinform. 2010, 78, 1266–1281. [Google Scholar] [CrossRef]
  20. Gomes, T.C.; Skaf, M.S. Cellulose-Builder: A toolkit for building crystalline structures of cellulose. J. Comput. Chem. 2012, 33, 1338–1346. [Google Scholar] [CrossRef]
  21. Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graph. 1996, 14, 33–38. [Google Scholar] [CrossRef]
  22. Phillips, J.C.; Hardy, D.J.; Maia, J.D.; Stone, J.E.; Ribeiro, J.V.; Bernardi, R.C.; Buch, R.; Fiorin, G.; Hénin, J.; Jiang, W.; et al. Scalable molecular dynamics on CPU and GPU architectures with NAMD. J. Chem. Phys. 2020, 153, 044130. [Google Scholar] [CrossRef] [PubMed]
  23. Hatcher, E.; Guvench, O.; MacKerell, A.D., Jr. CHARMM additive all-atom force field for aldopentofuranoses, methyl-aldopentofuranosides, and fructofuranose. J. Phys. Chem. B 2009, 113, 12466–12476. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Jorgensen, W.L.; Chandrasekhar, J.; Madura, J.D.; Impey, R.W.; Klein, M.L. Comparison of simple potential functions for simulating liquid water. J. Chem. Phys. 1983, 79, 926–935. [Google Scholar] [CrossRef]
  25. Harvey, M.; De Fabritiis, G. An implementation of the smooth particle mesh Ewald method on GPU hardware. J. Chem. Theory Comput. 2009, 5, 2371–2377. [Google Scholar] [CrossRef]
  26. Langan, P.; Nishiyama, Y.; Chanzy, H. X-ray structure of mercerized cellulose II at 1 Å resolution. Biomacromolecules 2001, 2, 410–416. [Google Scholar] [CrossRef]
  27. Souza, P.C.; Alessandri, R.; Barnoud, J.; Thallmair, S.; Faustino, I.; Grünewald, F.; Patmanidis, I.; Abdizadeh, H.; Bruininks, B.M.; Wassenaar, T.A.; et al. Martini 3: A general purpose force field for coarse-grained molecular dynamics. Nat. Methods 2021, 18, 382–388. [Google Scholar] [CrossRef]
  28. Souza, P.C.; Thallmair, S.; Marrink, S.J.; Mera-Adasme, R. An Allosteric Pathway in Copper, Zinc Superoxide Dismutase Unravels the Molecular Mechanism of the G93A Amyotrophic Lateral Sclerosis-Linked Mutation. J. Phys. Chem. Lett. 2019, 10, 7740–7744. [Google Scholar] [CrossRef]
  29. Liu, Z.; Moreira, R.A.; Dujmović, A.; Liu, H.; Yang, B.; Poma, A.B.; Nash, M.A. Mapping mechanostable pulling geometries of a therapeutic anticalin/CTLA-4 protein complex. Nano Lett. 2022, 22, 179–187. [Google Scholar] [CrossRef]
  30. Liaci, A.M.; Steigenberger, B.; de Souza, P.C.T.; Tamara, S.; Gröllers-Mulderij, M.; Ogrissek, P.; Marrink, S.J.; Scheltema, R.A.; Förster, F. Structure of the human signal peptidase complex reveals the determinants for signal peptide cleavage. Mol. Cell 2021, 81, 3934–3948. [Google Scholar] [CrossRef]
  31. Croom, F.H. Basic Concepts of Algebraic Topology; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2012. [Google Scholar]
  32. Abraham, M.J.; Murtola, T.; Schulz, R.; Páll, S.; Smith, J.C.; Hess, B.; Lindahl, E. GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 2015, 1, 19–25. [Google Scholar] [CrossRef] [Green Version]
  33. Tironi, I.G.; Sperb, R.; Smith, P.E.; van Gunsteren, W.F. A generalized reaction field method for molecular dynamics simulations. J. Chem. Phys. 1995, 102, 5451–5459. [Google Scholar] [CrossRef]
  34. Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling through velocity rescaling. J. Chem. Phys. 2007, 126, 014101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Parrinello, M.; Rahman, A. Crystal structure and pair potentials: A molecular-dynamics study. Phys. Rev. Lett. 1980, 45, 1196. [Google Scholar] [CrossRef]
  36. Poma, A.B.; Cieplak, M.; Theodorakis, P.E. Combining the MARTINI and structure-based coarse-grained approaches for the molecular dynamics studies of conformational transitions in proteins. J. Chem. Theory Comput. 2017, 13, 1366–1374. [Google Scholar] [CrossRef]
  37. Chwastyk, M.; Bernaola, A.P.; Cieplak, M. Statistical radii associated with amino acids to determine the contact map: Fixing the structure of a type I cohesin domain in the Clostridium thermocellum cellulosome. Phys. Biol. 2015, 12, 046002. [Google Scholar] [CrossRef] [Green Version]
  38. Okazaki, K.i.; Koga, N.; Takada, S.; Onuchic, J.N.; Wolynes, P.G. Multiple-basin energy landscapes for large-amplitude conformational motions of proteins: Structure-based molecular dynamics simulations. Proc. Natl. Acad. Sci. USA 2006, 103, 11844–11849. [Google Scholar] [CrossRef] [Green Version]
  39. Mahmood, M.I.; Poma, A.B.; Okazaki, K.i. Optimizing Gō-MARTINI coarse-grained model for F-BAR protein on lipid membrane. Front. Mol. Biosci. 2021, 8, 619381. [Google Scholar] [CrossRef]
Figure 1. Snapshots of the cross-section for the relaxed atomistic structures of crystalline cellulose fibrils: (a) I α , (b) I β , and (c) type-II, and and crystallographic unit cells are shown on the right side, and the values are taken from Refs. [9,10,26]. D-glucose molecule is shown in licorice representation with oxygens atoms in red and carbon atoms in cyan colours. Hydrogens were not included in this representation for sake of clarity. The chemical structure of cellulose fibrils were build by Cellulose-builder toolkit and rendered by VMD software.
Figure 1. Snapshots of the cross-section for the relaxed atomistic structures of crystalline cellulose fibrils: (a) I α , (b) I β , and (c) type-II, and and crystallographic unit cells are shown on the right side, and the values are taken from Refs. [9,10,26]. D-glucose molecule is shown in licorice representation with oxygens atoms in red and carbon atoms in cyan colours. Hydrogens were not included in this representation for sake of clarity. The chemical structure of cellulose fibrils were build by Cellulose-builder toolkit and rendered by VMD software.
Molecules 27 00976 g001
Figure 2. (a) Cellulose sheet representation of the CG model for the cellulose fibrils. Mapping of the atom, O2, O3, C6, and O6 into CG beads, CG1, CG2, CG3, and CG4, respectively, (top panel). Names of the heavy atoms in D-glucose are shown on the lower right. Solid lines indicate the bonded interactions of the model, dashed lines show the plane triangulation of a single cellulose chain, and thin dashed lines represent the non-bonded interaction between two parallel chains. (b) Equilibrated snapshot of two CG cellulose chains from two cellulose sheets for the I α allomorph. The intra-residue bonds are depicted using cyan and inter-residue bonds by grey and orange.
Figure 2. (a) Cellulose sheet representation of the CG model for the cellulose fibrils. Mapping of the atom, O2, O3, C6, and O6 into CG beads, CG1, CG2, CG3, and CG4, respectively, (top panel). Names of the heavy atoms in D-glucose are shown on the lower right. Solid lines indicate the bonded interactions of the model, dashed lines show the plane triangulation of a single cellulose chain, and thin dashed lines represent the non-bonded interaction between two parallel chains. (b) Equilibrated snapshot of two CG cellulose chains from two cellulose sheets for the I α allomorph. The intra-residue bonds are depicted using cyan and inter-residue bonds by grey and orange.
Molecules 27 00976 g002
Figure 3. All-atom MD distribution of distances between atomic centers considered as CG sites. The labels assigned to distances are described in the method section, as well as the relationship between first and second D-glucose neighbours denoted by “+” and “++” respectively. Data for cellulose I α , I β and type-II are represented by purple, green, and yellow solid lines, respectively. The horizontal black line indicates the value of the optimal distance parameter in the CG model. The distributions are bell-shaped with the exceptions of the intra-glucose distances O2-O6 and O3-O6, which follow a bimodal distribution.
Figure 3. All-atom MD distribution of distances between atomic centers considered as CG sites. The labels assigned to distances are described in the method section, as well as the relationship between first and second D-glucose neighbours denoted by “+” and “++” respectively. Data for cellulose I α , I β and type-II are represented by purple, green, and yellow solid lines, respectively. The horizontal black line indicates the value of the optimal distance parameter in the CG model. The distributions are bell-shaped with the exceptions of the intra-glucose distances O2-O6 and O3-O6, which follow a bimodal distribution.
Molecules 27 00976 g003
Figure 4. Time average RMSD of I α , I β , and type-II cellulose fibrils as a function of the Martini 3 bead types and three different elastic constants for O2-O6 and O3-O6 intramolecular distances. Reference snapshot for each system is the last frame from all-atom MD simulations. Some bead types are not shown in the plot as they induced large CG-MD instability during the simulation.
Figure 4. Time average RMSD of I α , I β , and type-II cellulose fibrils as a function of the Martini 3 bead types and three different elastic constants for O2-O6 and O3-O6 intramolecular distances. Reference snapshot for each system is the last frame from all-atom MD simulations. Some bead types are not shown in the plot as they induced large CG-MD instability during the simulation.
Molecules 27 00976 g004
Figure 5. Distribution of the fibril twist angle defined by the dihedral angle formed by O2 and C6 centers from two subsequent D-glucose. Panel (a) shows AA-MD representation and bottom panel (b) for the CG-MD case.
Figure 5. Distribution of the fibril twist angle defined by the dihedral angle formed by O2 and C6 centers from two subsequent D-glucose. Panel (a) shows AA-MD representation and bottom panel (b) for the CG-MD case.
Molecules 27 00976 g005
Table 1. Optimal CG bonded parameters for cellulose I allomorphs and type-II fibrils. Parameters are defined within a D-glucose and between the two subsequent neighbours.
Table 1. Optimal CG bonded parameters for cellulose I allomorphs and type-II fibrils. Parameters are defined within a D-glucose and between the two subsequent neighbours.
D-Glucose NeighboursCenter Pairs r i j (Å) k i j (kJ/mol)
noneO2-O32.8830,000
noneO2-C65.6230,000
noneO3-C64.9430,000
noneC6-O61.4330,000
noneO2-O66.422500
noneO3-O65.762500
1stO2-O27.5630,000
1stO3-O36.6930,000
1stC6-C67.4730,000
2ndO2-O210.4430,000
2ndO3-O310.4430,000
2ndC6-C610.4430,000
Table 2. Optimal Martini 3 nonbonded parameters for cellulose I allomorphs (I α and I β ) and type-II. CG3 is defined by bead type TC1 and CG1, CG2 and CG4 as SP6 bead type.
Table 2. Optimal Martini 3 nonbonded parameters for cellulose I allomorphs (I α and I β ) and type-II. CG3 is defined by bead type TC1 and CG1, CG2 and CG4 as SP6 bead type.
Interaction between CG Bead Types σ i j (Å) ϵ i j (kJ/mol)
WSP64.2504.530
SP6SP64.1004.290
WTC14.1500.550
TC1TC13.4001.510
SP6TC14.8400.890
Table 3. Comparison of the distances between CG centres between AA-MD and CG-MD models for I α , I β , and type II fibrils. Mean value and standard deviation are shown for each entry. The first column indicates whether the distance is found inside the D-glucose (none) or between the two subsequent (1st and 2nd neighbours) D-glucose molecules.
Table 3. Comparison of the distances between CG centres between AA-MD and CG-MD models for I α , I β , and type II fibrils. Mean value and standard deviation are shown for each entry. The first column indicates whether the distance is found inside the D-glucose (none) or between the two subsequent (1st and 2nd neighbours) D-glucose molecules.
D-glucose NeighboursPairsFibrilDistances (Å)
AA-MDCG-MD
SP6/2500
NoneO2-O3I α 2.87 ± 0.282.88 ± 0.23
NoneO2-O3I β 2.88 ± 0.282.88 ± 0.23
NoneO2-O3II2.91 ± 0.282.88 ± 0.23
NoneO2-C6I α 5.62 ± 0.275.63 ± 0.25
NoneO2-C6I β 5.63 ± 0.275.63 ± 0.25
NoneO2-C6II5.64 ± 0.275.62 ± 0.25
NoneO3-C6I α 4.96 ± 0.284.94 ± 0.25
NoneO3-C6I β 4.95 ± 0.284.93 ± 0.25
NoneO3-C6II4.93 ± 0.284.94 ± 0.25
NoneC6-O6I α 1.43 ± 0.161.43 ± 0.27
NoneC6-O6I β 1.43 ± 0.161.42 ± 0.27
NoneC6-O6II1.43 ± 0.161.42 ± 0.27
NoneO2-O6I α 6.47 ± 0.426.29 ± 0.57
NoneO2-O6I β 6.38 ± 0.466.31 ± 0.57
NoneO2-O6II6.35 ± 0.356.57 ± 0.49
NoneO3-O6I α 5.69 ± 0.585.65 ± 0.56
NoneO3-O6I β 5.62 ± 0.605.64 ± 0.56
NoneO3-O6II6.08 ± 0.475.49 ± 0.50
1stO2-O2I α 7.57 ± 0.357.56 ± 0.25
1stO2-O2I β 7.57 ± 0.347.56 ± 0.25
1stO2-O2II7.58 ± 0.357.56 ± 0.25
1stO3-O3I α 6.69 ± 0.346.70 ± 0.24
1stO3-O3I β 6.70 ± 0.346.71 ± 0.24
1stO3-O3II6.70 ± 0.356.70 ± 0.24
1stC6-C6I α 7.48 ± 0.387.46 ± 0.26
1stC6-C6I β 7.47 ± 0.377.47 ± 0.26
1stC6-C6II7.46 ± 0.367.47 ± 0.26
2ndO2-O2I α 10.44 ± 0.4510.44 ± 0.24
2ndO2-O2I β 10.44 ± 0.4410.43 ± 0.24
2ndO2-O2II10.42 ± 0.4510.44 ± 0.24
2ndO3-O3I α 10.44 ± 0.4010.43 ± 0.24
2ndO3-O3I β 10.44 ± 0.3910.43 ± 0.24
2ndO3-O3II10.42 ± 0.4010.44 ± 0.24
2ndC6-C6I α 10.44 ± 0.4610.44 ± 0.26
2ndC6-C6I β 10.44 ± 0.4610.44 ± 0.26
2ndC6-C6II10.42 ± 0.4510.44 ± 0.26
Table 4. RMSD cross-system for all cellulose fibril trajectories and their respective standard deviations.
Table 4. RMSD cross-system for all cellulose fibril trajectories and their respective standard deviations.
AA-MDCG-MDTrajectory RMSD(Å)
I α I α 4.264 ± 0.751
I α I β 16.860 ± 0.127
I α II21.111 ± 0.235
I β I α 16.867 ± 0.120
I β I β 5.440 ± 0.681
I β II18.786 ± 0.390
III α 20.290 ± 0.091
III β 17.840 ± 0.130
IIII3.740 ± 0.400
Table 5. Statistics of twist angles for all cellulose fibrils. The error in the measurements corresponds to the standard deviation (SD) for each entry.
Table 5. Statistics of twist angles for all cellulose fibrils. The error in the measurements corresponds to the standard deviation (SD) for each entry.
AverageSDMaxMin
AA-MD
I α 3.55758711.986272−136.20123.50
I β 2.5507897.222558−167.0099.80
II−0.4209026.886370−150.00173.40
CG-MD
I α 1.4061226.499510−54.1061.10
I β 1.2513996.292622−33.8089.30
II−0.4279399.148864−107.20120.20
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Moreira, R.A.; Weber, S.A.L.; Poma, A.B. Martini 3 Model of Cellulose Microfibrils: On the Route to Capture Large Conformational Changes of Polysaccharides. Molecules 2022, 27, 976. https://doi.org/10.3390/molecules27030976

AMA Style

Moreira RA, Weber SAL, Poma AB. Martini 3 Model of Cellulose Microfibrils: On the Route to Capture Large Conformational Changes of Polysaccharides. Molecules. 2022; 27(3):976. https://doi.org/10.3390/molecules27030976

Chicago/Turabian Style

Moreira, Rodrigo A., Stefan A. L. Weber, and Adolfo B. Poma. 2022. "Martini 3 Model of Cellulose Microfibrils: On the Route to Capture Large Conformational Changes of Polysaccharides" Molecules 27, no. 3: 976. https://doi.org/10.3390/molecules27030976

Article Metrics

Back to TopTop