ACS Publications. Most Trusted. Most Cited. Most Read
My Activity
CONTENT TYPES

Figure 1Loading Img
RETURN TO ISSUEPREVB: Biophysical and B...B: Biophysical and Biochemical Systems and ProcessesNEXT

Computational Analysis on the Allostery of Tryptophan Synthase: Relationship between α/β-Ligand Binding and Distal Domain Closure

  • Shingo Ito
    Shingo Ito
    Theoretical Molecular Science Laboratory, RIKEN Cluster for Pioneering Research, 2-1 Hirosawa, Wako, Saitama 351-0198, Japan
    More by Shingo Ito
  • Kiyoshi Yagi
    Kiyoshi Yagi
    Theoretical Molecular Science Laboratory, RIKEN Cluster for Pioneering Research, 2-1 Hirosawa, Wako, Saitama 351-0198, Japan
    More by Kiyoshi Yagi
  • , and 
  • Yuji Sugita*
    Yuji Sugita
    Theoretical Molecular Science Laboratory, RIKEN Cluster for Pioneering Research, 2-1 Hirosawa, Wako, Saitama 351-0198, Japan
    Computational Biophysics Research Team, RIKEN Center for Computational Science, 7-1-26 Minatojima-Minamimachi, Chuo-ku, Kobe, Hyogo 650-0047, Japan
    Laboratory for Biomolecular Function Simulation, RIKEN Center for Biosystems Dynamics Research, 1-6-5 Minatojima-Minamimachi, Chuo-ku, Kobe, Hyogo 650-0047, Japan
    *Email: [email protected]
    More by Yuji Sugita
Cite this: J. Phys. Chem. B 2022, 126, 17, 3300–3308
Publication Date (Web):April 21, 2022
https://doi.org/10.1021/acs.jpcb.2c01556

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0.
  • Open Access

Article Views

827

Altmetric

-

Citations

3
LEARN ABOUT THESE METRICS
PDF (7 MB)
Supporting Info (2)»

Abstract

Tryptophan synthase (TRPS) is a bifunctional enzyme consisting of α and β-subunits and catalyzes the last two steps of l-tryptophan (L-Trp) biosynthesis, namely, cleavage of 3-indole-d-glycerol-3′-phosphate (IGP) into indole and glyceraldehyde-3-phosphate (G3P) in the α-subunit, and a pyridoxal phosphate (PLP)-dependent reaction of indole and l-serine (L-Ser) to produce L-Trp in the β-subunit. Importantly, the IGP binding at the α-subunit affects the β-subunit conformation and its ligand-binding affinity, which, in turn, enhances the enzymatic reaction at the α-subunit. The intersubunit communications in TRPS have been investigated extensively for decades because of the fundamental and pharmaceutical importance, while it is still difficult to answer how TRPS allostery is regulated at the atomic detail. Here, we investigate the allosteric regulation of TRPS by all-atom classical molecular dynamics (MD) simulations and analyze the potential of mean-force (PMF) along conformational changes of the α- and β-subunits. The present simulation has revealed a widely opened conformation of the β-subunit, which provides a pathway for L-Ser to enter into the β-active site. The IGP binding closes the α-subunit and induces a wide opening of the β-subunit, thereby enhancing the binding affinity of L-Ser to the β-subunit. Structural analyses have identified critical hydrogen bonds (HBs) at the interface of the two subunits (αG181-βS178, αP57-βR175, etc.) and HBs between the β-subunit (βT110 – βH115) and a complex of PLP and L-Ser (an α-aminoacrylate intermediate). The former HBs regulate the allosteric, β-subunit opening, whereas the latter HBs are essential for closing the β-subunit in a later step. The proposed mechanism for how the interdomain communication in TRPS is realized with ligand bindings is consistent with the previous experimental data, giving a general idea to interpret the allosteric regulations in multidomain proteins.

This publication is licensed under

CC-BY-NC-ND 4.0.
  • cc licence
  • by licence
  • nc licence
  • nd licence

1. Introduction

ARTICLE SECTIONS
Jump To

The tryptophan (Trp) biosynthesis has recently drawn great attention as a potential target for anti-tuberculosis (TB) drugs. (1,2) TB disease is one of the globally spread epidemics, where 1.3 million deaths have been estimated in 2021. (3) Furthermore, the treatment of multidrug-resistant, pathogenic bacteria, Mycobacterium tuberculosis (M.tb), requires an urgent need for new therapeutic approaches with a novel molecular target. Tryptophan synthase (TRPS), which catalyzes the last two steps of Trp biosynthesis, is a bifunctional enzyme consisting of two subunits, α and β. One of the unique features of TRPS is an allosteric regulation, where conformational changes of the α- and β-subunits are synchronized with the activity of enzymatic reactions. (4,5) Recent studies have reported effective allosteric inhibitors of TRPS that kills M.tb. (6,7) Nonetheless, further understanding in an atomic detail is necessary for precise control of the allosteric inhibition.
The structure and reaction of TRPS from Salmonella typhimurium have been extensively studied since the 1980s. (8−10) The two subunits of TRPS have their own active sites, which are called the α- and β-active sites. At the α-active site, the α-reaction cleaves 3-indole-d-glycerol-3′-phosphate (IGP) into indole and glyceraldehyde-3-phosphate (G3P). The indole is transferred to the β-active site through an internal tunnel of 25 Å length, preventing the leak of indole during the reaction cycle. At the β-active site, the β-reaction proceeds in two stages. First, the pyridoxal phosphate (PLP) co-factor, covalently bound to βK87 of the β-subunit [denoted E(Ain)], reacts with a ligand, l-serine (L-Ser), to produce an α-aminoacrylate [denoted E(A-A)]. In the second stage, E(A-A) reacts with the indole transferred from the α-active site to yield the final product, L-Trp. Interestingly, the α and β-reactions as well as the indole transfer are functionally coupled to each other. (1) The binding of IGP at the α-active site enhances the affinity of L-Ser binding at the β-active site (11) and promotes the reaction of L-Ser and PLP to produce E(A-A). (12,13) (2) The formation of E(A-A) promotes the cleavage of IGP, the indole transfer, and the production of L-Trp. (12,13) The whole scheme of the Trp synthesis is schematically drawn in Figure S1. The allosteric regulation has been extensively investigated by structural and kinetic studies combined with site-directed mutagenesis. (14−24) These studies have revealed two important regions, a flexible loop 6 in the α-subunit (αL6) and a communication (COMM) domain in the β-subunit, which are located in the proximity of the two active sites as shown in Figure 1A. These regions not only serve as gates of the α- and β-active sites but also interact with each other at the interface of the two subunits. The hydrogen bond (HB) between αG181 in αL6 and βS178 in the COMM domain has been suggested to be critical for the ligand-induced conformational changes. (25,26)

Figure 1

Figure 1. (A) Essential domains related to the allostery of TRPS located at the interface of α and β-subunits: αL2 (purple, residue ID 53–60), αL6 (red, residue ID 176–192), αH2 (cyan, residue ID 61–75), and the COMM domain (blue, residue ID 102–189). (B) Entry pathway proposed from the X-ray crystal structures to the β-active site (green). (C) Bottleneck along the proposed L-Ser entry pathway. The white and blue vdW spheres represent the β-subunit and the COMM domain, respectively. The surface of the L-Ser (yellow colored wire) overlaps with the surface of TRPS (vdW sphere). Detailed information on how to set L-Ser in the pathway is described in the Supporting Information.

Despite accumulated structural and functional information of TRPS, there are still many questions to be answered regarding the allostery of TRPS. One of the most important questions is how IGP binding at the α-active site can enhance the ligand affinity at the distal β-active site. The entry pathway proposed from the X-ray crystal structures is too narrow for the L-Ser entry either in the Apo|E(Ain) state (PDB ID: 1K8X (20)) (Figure 1B,C) or in the IGP|E(Ain) state (PDB ID: 1WBJ (21)), suggesting the importance of dynamic structures of TRPS. Hereafter, we use the notation A|B, where A and B indicate the bound (and reaction) states of the α- and β-active site, respectively. After the β-subunit is widely opened for making sufficient space for the L-Ser entry, it should be closed for enhancing the reaction at the β-active site. Mechanisms for this domain closure also remain elusive. To answer these questions, dynamic structures of TRPS and interdomain interaction in different ligand-bound states are necessary in addition to the static X-ray crystal structures.
Molecular dynamics (MD) simulations of biomacromolecules have made great progress in the last decade. (27−29) Developments of computational algorithms and MD-specialized supercomputers (30) have made feasible the simulation in a time scale of 100 μs or even milliseconds. Enhanced sampling methods are another way to enable an efficient sampling of conformational space and the computation of free-energy landscapes of biomolecules. (31−33) Replica-exchange molecular dynamics (REMD), (34−36) generalized replica exchange solute tempering (gREST), (37,38) Gaussian accelerated MD (GaMD), (39) and metadynamics (40) are representative algorithms in enhanced sampling algorithms of biomolecular simulations. They are ready to use in most of MD software with/without additional tools, such as PLUMED. (41) In our MD program GENESIS (https://www.r-ccs.riken.jp/labs/cbrt/), (42,43) many enhanced sampling algorithms are available for users just by preparing input scripts.
Several important computational studies on TRPS have been reported in the literature. (44−50) In a pioneering work by Fatmi et al., (45) all-atom MD simulations of TRPS were performed in the apo and ligand-bound [with IGP and E(A-A)] states to clarify ligand-induced conformational changes of αL6 and the COMM domain, although the sampling was limited to a 60 ns conventional MD in those days. More recently, Maria-Solano et al. (48) studied the TRPS from Pyrococcus furiosus (Pf TrpS) by metadynamics simulations. The potential of mean force (PMF) was computed along the open-to-closed conformational changes of the COMM domain in various intermediates of the β-reaction to discuss the stability of Pf TrpS (α- and β-complex) and a stand-alone β-subunit with mutations. To the best of our knowledge, there are no computational studies on the allosteric regulation of the β-subunit motions induced by IGP binding at the α-subunit.
To reveal the molecular mechanisms for the allosteric structures of TRPS, we have carried out extensive MD simulations of TRPS in three different ligand-bound states, namely, Apo|E(Ain), IGP|E(Ain), and IGP|E(A-A) bound states. The PMF in each state was computed using the umbrella sampling (US) method. (51) The main finding in the simulation is that the COMM domain takes a stable, widely open form upon the IGP binding at the α-binding site, and that the widely open form provides an entry pathway for L-Ser with a sufficient space to reach the β-active site. Structural analyses of the simulation trajectories show that the IGP binding induces a closure of αL6 and the formation of HBs between loop 2, helix 2 of the α-subunit (αL2 and αH2, respectively, see Figure 1A), and the COMM domain. The COMM domain motions for opening the L-Ser entry pathway take place subsequently. We also show that the E(A-A) in the β-active site forms strong HBs with a portion of the COMM domain (βT110−βH115), which stabilizes a closed conformation of the COMM domain.

2. Computational Details

ARTICLE SECTIONS
Jump To

2.1. System Setup

All-atom models of three different physiological states, namely, Apo|E(Ain), IGP|E(Ain), and IGP|E(A-A) states were prepared based on the X-ray crystal structures; PDBID: 1K8X, (20) 1WBJ, (21) and 2J9X, (13) respectively. In the IGP bound states, the inhibitor in the α-active site was replaced manually with IGP in reference to the crystal structure, PDBID: 1QOQ. (22) Cs+ in the monovalent cation site of 2J9X was replaced with Na+. The atomic structures of missing residues were modeled by GalaxyFill (52) via CHARMM-GUI. (53) αI87 of 1K8X and 1WBJ was replaced to αL87 to be consistent with 2J9X. The information of all-atom models used in MD simulations is summarized in Table 1. Topology files used in the MD simulations were generated by antechamber and tleap module of AMBER18. (27) The force field was set to FF14SB (54) for TRPS and generalized AMBER force field 2 (GAFF2) (55,56) for the ligands. The charges of the ligands were parametrized by quantum chemical calculations at the AM1 level. TRPS in each state was set in a cubic box of 120 × 120 × 120 Å3 with TIP3P water molecules, and the system was neutralized by 150 mM K+ and Cl ions. All molecular graphic images were generated using Visual Molecular Dynamics (VMD) (57) and PyMOL. (58)
Table 1. Model of Apo|E(Ain), IGP|E(Ain), and IGP|E(A-A) States of TRPS Used in the MD Simulations
state PDB α active site β active site modeled residues protonated residues initial domain structure (α |β)
Apo|E(Ain) 1K8X20 none E(Ain) α: 178–195, 268 αE49 open|open
β: 1, 393–397
IGP|E(Ain) 1WBJ21 IGP-bound E(Ain) α: 268 αE49 closed|open
β: 392–397
IGP|E(A-A) 2J9X13 IGP-bound E(A-A) α: 1, 190–192 αE49, βK87 closed|closed
β: 397

2.2. MD Simulations

Each simulation system was first optimized with energy minimization and then equilibrated by MD simulations in several steps. The details of the equilibration steps are summarized in Tables S1 and S2. MD simulations in the NVT and NPT ensembles were carried out for 4.6 ns in the first seven equilibration steps followed by those with umbrella potentials for 11 ns in two steps. Then, the production runs of US-MD simulations were performed for 50 ns with 22 umbrella windows (1.1 μs in total) in the NVT ensemble. The temperature and pressure were controlled at 300 K and 1 atm using the BUSSI thermostat. (59) The time step was set to 2.0 fs. Reference system propagator algorithm (RESPA) (60) was used to calculate the reciprocal-space interactions and forces every other step. The smooth particle-mesh Ewald (PME) method (61,62) was used to calculate long-range electrostatic interactions, and the cutoff distance was set as 8.0 Å. The reaction coordinate of US-MD was defined in terms of distances of a salt bridge (SB) between βR141 and βD305 and a HB between βR141 and βS299
(1)
(2)
(3)
The reference values of ξCV1, ξCV2, and ξCV3 for each window are shown in Table S3. The force constant of the umbrella potential was set to 10.0 kcal mol–1 Å–2. The probability distributions along ξCV3 show good overlaps between neighboring histograms (see Figure S2). All bonds involving hydrogen atoms were constrained using the SHAKE/RATTLE (63) and SETTLE (64) algorithms. The GENESIS 2.0β program (65) was used in all MD simulations.

2.3. Analyses

The trajectory data of the last 30 ns of the production run were used to compute the PMFs and the HB probabilities in the Apo|E(Ain), IGP|E(Ain), and IGP|E(A-A) states. PMF was calculated using the MBAR (66) method implemented in GENESIS analysis tools. The probability of HBs was calculated by counting the number of HBs between αL2, αL6, αH2, and the COMM domain for each snapshot of the trajectory and weight-averaging the count number using the weight factor of each snapshot from MBAR. The HB was defined by the heavy atom distance, X–(H)–Y (X and Y represent oxygen or nitrogen), being smaller than 3.6 Å.
Tunnels from the β-active site to the surface of TRPS were analyzed using CAVER 3.0 software (67) in the Apo|E(Ain), IGP|E(Ain), and IGP|E(A-A) states after the US-MD simulations. The atom of the contact point in E(A-A), i.e., C4A, was set to a starting point of the tunnel analysis. Other parameters were set to probe_radius = 1.2 or 1.8 Å, shell_radius = 3.5 Å, shell_depth = 4.0 Å, and clustering_threshold = 12.0 Å.

3. Results

ARTICLE SECTIONS
Jump To

3.1. Conformational Changes of the COMM Domain upon Ligand Binding

To examine the motion of the COMM domain upon the α- and β-ligand bindings, we calculated two-dimensional (2D-) PMFs as a function of ξCV3 (eq 3) and the root mean square deviation (RMSD) of the COMM domain. The latter was measured by aligning the β-subunit except for the COMM domain in reference to an X-ray crystal structure, PDB ID: 2J9X (IGP|E(A-A) state). X-ray crystal structures, in which the COMM domain is closed (PDB ID: 3CEP, 3PR2) and open (PDB ID: 1K8X, 1WBJ), give the RMSD value of 0.58, 0.70, 2.74, and 2.67 Å, respectively. Thus, the COMM domain is referred to as open conformation when the RMSD is larger than 2.0 Å. The resulting 2D-PMFs are shown in Figure 2. In the Apo|E(Ain) state, stable structures mainly exist around ξCV3 = 16 Å and RMSD = 2.0–3.0 Å (Figure 2A), suggesting that the COMM domain takes an open form and can hardly access a closed form. In contrast, in the IGP|E(Ain) state, the stable structure is shifted toward a region of ξCV3 = 20–23 Å and RMSD = 2.8–3.8 Å (Figure 2B, yellow circle), an intriguing structure newly predicted in the present MD simulation. We emphasize that the COMM domain is more widely open than that of the X-ray crystal structure of the IGP|E(Ain) state (PDB ID: 1WBJ).

Figure 2

Figure 2. 2D-PMFs along ξCV3 (eq 3) and the RMSD of the COMM domain in reference to an X-ray crystal structure (PDB ID: 2J9X (13)) in the Apo|E(Ain) (A), IGP|E(Ain) (B), and IGP|E(A-A) (C) states. 1D-PMFs along ξCV3 and RMSD are also shown in the bottom and left of each panel, respectively. X-ray crystal structures of Apo|E(Ain) (PDB ID: 1K8X (20)), IGP|E(Ain) (PDB ID: 1WBJ (21)), IGP|E(A-A) (PDB ID: 2J9X, 3CEP, (68) and 3PR2 (69)) states are shown in red, green, and blue circles, respectively. The yellow circle in the IGP|E(Ain) state represents a widely open COMM domain obtained in the present MD simulation.

To further characterize the widely open form of the IGP|E(Ain), we counted the number of cavities connecting the β-active site and the surface of β-subunit, i.e., tunnels for L-Ser to access the β-active site. The number of tunnels averaged over 10 snapshot structures is listed in Table 2 (the data of each snapshot are shown in Tables S4–S6). In the closed form of the IGP|E(A-A) state, no such tunnel is detected irrespective of the size of the probe radius. On the other hand, at least one tunnel is detected in the open form of both Apo|E(Ain) and IGP|E(Ain) states with a probe radius (rp) of 1.2 Å. However, the average number of tunnels with rp = 1.8 Å is reduced to 0.2 in the open form of the Apo|E(Ain) state, whereas it is kept to 0.7 in the widely open form of the IGP|E(Ain) state. The result suggests that the tunnel is wider in the IGP|E(Ain) state than in the Apo|E(Ain) state, which is consistent with the RMSD value of 3.13 and 2.20 Å, respectively. Note that the tunnel with rp = 1.8 Å is absent in the X-ray crystal structures, 1K8X and 1WBJ, which are both in the open form. The entry pathway for L-Ser in the widely open form is visualized in Figure 3. The entrance is located in the vicinity of the SB pair, βD305/βR141, and its size is sufficiently large to allow L-Ser to reach the β-active site.

Figure 3

Figure 3. (A) Tunnel from the β-active site to the surface of TRPS (green) of the widely open form in the MD simulations of the IGP|E(Ain) state detected by CAVER 3.0. (67) βD305 and βR141 are shown in red, where the SB is broken in this case. (B) Smallest bottleneck radius was larger than the surface of the L-Ser (yellow wired surface). Detailed information on how to set the L-Ser in the pathway is described in the Supporting Information.

Table 2. Average Number of Tunnels from the β-Active Site to the Surface of β-Subunit (Ntun) Obtained from 10 Snapshot Structures of Each State and X-ray Crystal Structures Using a Probe Radius (rp) of 1.2 and 1.8 Å
  ξCV3 RMSD [Å] Ntun with rp = 1.2 Ntun with rp = 1.8
MD (Apo|E(Ain)) 16.1 2.20 1.1 0.2
MD (IGP|E(Ain)) 22.4 3.13 1.1 0.7
MD (IGP|E(A-A)) 18.5 1.73 0.2 0.0
X-ray (1K8X) 15.7 2.74 1 0
X-ray (1WBJ) 16.0 2.67 1 0
X-ray (2J9X) 5.89 0.00 1 0
Figure 2C shows that the COMM domain is closed in the IGP|E(A-A) state. It is noticeable that the closure of the COMM domain is uncorrelated with the formation of the SB and HB. This is clearly seen in the 1D-PMF along ξCV3, where the free-energy minima are found not only around the SB/HB formation (ξCV3 = 6 Å) but also around ξCV3 = 19 Å. The free-energy barriers in this range are less than 1.0 kcal mol–1, suggesting that the SB/HB can easily form and break while the COMM domain is kept closed.
The result so far suggests two important features. One is that the IGP binding at the α-active site induces a conformational change of the COMM domain from an open form to a widely open form. Since the α-active site and the COMM domain are spatially remote, further elucidation of the mechanism for such an allosteric regulation is intriguing. Secondly, the COMM domain is closed in the IGP|E(A-A) state, yet the formation of the SB and HB is not mandatory. Now, the question is what is the essential interaction that leads to the closure of the COMM domain? We investigate these points in the following subsections.

3.2. Wide Opening of the COMM Domain

We calculated 2D-PMFs to understand the interaction between the αL6 and the COMM domain. The first CV was set to a distance between a nitrogen atom of αG181 in αL6 and an oxygen atom of βS178, which belongs to the COMM domain,
(4)
The previous structural studies have reported that the HB was formed between αG181 and βS178 in a closed form. (26) Therefore, we use this HB as a measure of the α-subunit conformation: when ζ < 3.75 Å the α-subunit is referred to as a closed form. The second CV was set to the RMSD of the COMM domain, the same as in Figure 2. The result is shown in Figure 4. In the Apo|E(Ain) state, both αL6 and COMM domains are in an open form sampling a wide configurational space (Figure 4A). The stable structure is found in a region around ζ = 12 ∼ 18 Å and RMSD = 2.0 ∼ 3.5 Å. Note that this region contains the X-ray crystal structure of the Apo|E(Ain) state, 1K8X.

Figure 4

Figure 4. Two-dimensional PMFs along ζ (eq 4) and the RMSD of a COMM domain in reference to 2J9X in the Apo|E(Ain) (A), IGP|E(Ain) (B), and IGP|E(A-A) (C) states. 1D-PMFs along ζ and RMSD are also shown in the bottom and left of each panel, respectively. X-ray crystal structures of Apo|E(Ain) (PDB ID: 1K8X), IGP|E(Ain) (PDB ID: 1WBJ), IGP|E(A-A) (PDB ID: 2J9X, 3CEP, and 3PR2) states are shown in red, green, and blue circles, respectively.

The IGP binding drastically changes the appearance of the 2D-PMF due to a closure of the αL6 (Figure 4B). The conformational change of αL6 is clearly observed in a 1D-PMF along ζ, which gives a deep minimum around ζ = 3.0. On the other hand, the COMM domain remains in an open form but also shifts the population toward a widely open form around RMSD ∼3.5 Å. The result suggests an allosteric communication between the α- and β-subunits, where the closure of αL6 stimulates the wide opening of the COMM domain. αL6 and COMM domains are both closed in the IGP|E(A-A) bound state (Figure 4C).
To elucidate how the α-subunit invokes the signal of IGP bindings, we focus on the change in HBs between αL2, αL6, αH2, and the COMM domain. We analyzed the probability of HB formation (pHB) in the last 30 ns trajectories of the Apo|E(Ain) and IGP|E(Ain) states. The results in Figure 5 show that the IGP binding triggers a growth of the HB network. The number of HBs with pHB > 0.3 is four in the Apo|E(Ain) state, whereas the number increases to six in the IGP|E(Ain) state. Furthermore, four of them are found to be strongly bound with pHB > 0.6 (highlighted in red). Two of the strong HBs, αG61-αT183 and αG181-βS178, are visualized in Figure 6A–C. The IGP forms a HB with αD60, which induces the formation of HBs between αG61−αT183 and αG181−βS178 and the change of αL6 conformation to a closed form. Other HBs, αD60-αQ65, αQ65-βS161, and αP57-βR175, are visualized in Figure 6D,E. The HB between αD60-αQ65 changes the conformation of the αH2 (cyan allow), which recruits other HBs and shifts the COMM domain to a widely open form (blue allow). Therefore, the HB network plays an essential role in the transfer of the signal of IGP binding to the β-subunit.

Figure 5

Figure 5. Probability of HB formation (left) and schematic representations of the HB network (right) in the Apo|E(Ain) and IGP|E(Ain) state. The main HBs with the probability >0.6 are shown in red color. Asterisk indicates the HBs present in the X-ray crystal structures of the IGP|E(Ain) state, 1WBJ.

Figure 6

Figure 6. HB network between αL2, αH2, αL6, and the COMM domain. (A–C) HBs between IGP-αD60, αG61-αT183, and αG181-βS178, which are present in the IGP|E(Ain) state (right) and absent in the Apo|E(Ain) state (left). (D, E) HBs between αD60-αQ65, αQ65-βS161, and αP57-βR175 formed in the IGP|E(Ain) state. The structure of the Apo|E(Ain) state is superimposed with a transparent gray color.

Although there are mutational studies on αG181−βS178, (25,26) no experimental study on mutation of αP57 or βR175 has been reported. To reinforce the importance of the intersubunit interaction between αP57−βR175, we mutated βR175 to βA175 and performed US-MD simulations in the same way as in the wild type (WT). The resulting 1D-PMFs along the RMSD of the COMM domain are shown in Figure 7. In the Apo|E(Ain) state, the PMF of the mutant is very similar to that of WT, indicating that the conformation of the Apo|E(Ain) state is unaffected by the mutation. In contrast, the PMF of the IGP|E(Ain) state of the mutant is different from that of the WT. The conformational change from open to widely open forms is suppressed in the mutant due to the lack of a HB between αP57−βR175.

Figure 7

Figure 7. 1D-PMF of RMSD of COMM domain in the WT–Apo|E(Ain) (top-left), βR175A–Apo|E(Ain) (top right), WT– IGP|E(Ain) (bottom left), and βR175A– IGP|E(Ain) state (bottom right).

3.3. Closure of the COMM Domain

To clarify the driving force to close the COMM domain, we analyzed the HBs between the β-subunit and the β-ligand [E(Ain) or E(A-A)] in the last 30 ns trajectories of the IGP|E(Ain) and IGP|E(A-A) states. The probability of HB formation and the structure are shown in Figure 8. Regardless of the presence of L-Ser, the phosphate group of PLP is always hydrogen-bonded with βG232-βN236. These HBs stabilize the position of PLP in the β-active site. The difference in the HB network between IGP|E(Ain) and IGP|E(A-A) states is observed in a portion of the COMM domain, βT110-βH115, which is strongly hydrogen-bonded with an aminoacrylate moiety of E(A-A). As shown in Figure 8D, these HBs largely shift the position of βT110-βH115. Note that the intersubunit HBs are kept unchanged by the closure of the COMM domain (Figure S3). These results suggest that the closed form of the COMM domain is stabilized by strong HBs formed with the E(A-A) intermediate.

Figure 8

Figure 8. (A) Probability of HB formation in the IGP|E(Ain) and IGP|E(A-A) states. AA and PP indicate the α-aminoacrylate and phosphate group, respectively. Asterisk represents the HBs present in the X-ray crystal structure, 2J9X. (B, C) HBs between the β-subunit and the β-ligand in the IGP|E(Ain) and IGP|E(A-A) states. (D) Superposition of the conformation in the IGP|E(Ain) state (gray transparent) and IGP|E(A-A) state (colored) around the β-ligand. The two structures were superimposed by all Cα atoms of TRPS except for those of αL2, αL6, αH2, and COMM.

The remaining question is when is the COMM domain closed, i.e., upon the binding of L-Ser to the β-active site or the formation of E(A-A) after the reaction of PLP and L-Ser? To clarify this point, we have carried out an additional MD simulation after converting E(A-A) to E(Ain) and adding L-Ser [denoted E(Ain).L-Ser]. The initial structure was taken from one of the stable structures of IGP|E(A-A). After the system was equilibrated as before (see Table S1), the production run was performed for 50 ns. Figure 9 shows the time series of RMSD of the COMM domain along the simulation time. Although the RMSD is slightly increased in the initial stage, it is stabilized around 2.0 Å after 30 ns. The COMM domain does not reach a widely opened form, and L-Ser remains intact at the β-active site. The probability of HB formation and the structure of IGP|E(Ain).L-Ser are shown in Figure 10. Although the HBs with βG113-βH115 are absent, L-Ser is strongly hydrogen-bonded with βG110-βH112. It is plausible that these HBs prevent wide opening of the COMM domain. Therefore, we suggest that the L-Ser binding induces the conformational change of the COMM domain from widely open to open, and the COMM domain is further closed upon the formation of E(A-A).

Figure 9

Figure 9. RMSD of the COMM domain in the IGP|E(Ain).L-Ser state.

Figure 10

Figure 10. Probability of HB formation (A) and the structure (B) in the IGP|E(Ain).L-Ser state.

4. Discussion and Conclusions

ARTICLE SECTIONS
Jump To

MD simulations of TRPS have been carried out in three ligand bound states, namely, Apo|E(Ain), IGP|E(Ain), and IGP|E(A-A) states to elucidate the mechanism for the ligand-induced conformational changes of α and β-subunits. The PMF obtained from US simulations has revealed stable, widely opened conformations of the COMM domain in the IGP|E(Ain) state. The widely opened conformation has a sufficient space for L-Ser to enter into the β-active site, unlike the open conformation observed in X-ray crystal structures and MD simulations in Apo|E(Ain) state. Such a conformational change is consistent with a kinetic experiment, which reported an increase in the binding affinity of L-Ser upon the IGP binding at the α-subunit. (11) In the X-ray crystal structure, the pathway for L-Ser to access the β-active site is unclear. The position/orientation of the COMM domain of the IGP|E(Ain) state (PDBID: 1WBJ) exhibited little change from those in the Apo|E(Ain) state (PDBID: 1K8X). In another crystal structure (PDBID: 1QOQ), the COMM domain was reported to be widely opened. This structure was missing the essential monovalent cation at the β-subunit and therefore the biological meanings were not clear. The present result has clarified the entry pathway for L-Ser toward the β-active site, which passes through the SB pair (βD305 and βR141) to the β-active site (see Figure 3).
In the IGP|E(A-A) state, the α- and β-subunits are both found to be in a closed form. Although the SB between βD305 and βR141 and HB between βS299 and βR141 were suggested to play key roles in stabilizing the closed conformation of the COMM domain, (4) the present results showed that the formation of the SB and HB is uncorrelated with the closure of the COMM domain. Instead, the present calculation suggests that the closure is driven by strong HBs between an aminoacrylate moiety of the E(A-A) intermediate and a portion of the COMM domain (βT110-βH115). The finding is in line with the very recent mutagenesis experiment of βQ114A, which observed an alternation of the stability of the open and closed conformations of the COMM domain in the mutant. (24)
In summary, the present MD simulation has revealed the mechanism for allosteric regulation of TRPS in atomic details. (1) The IGP binding to the α-active site induces a closure of αL6, (2) the fix of αL6 stimulates the development of a HB network among αL2, αH2, αL6, and the COMM domain, (3) the intersubunit HBs, αG181−βS178 and αP57−βR175, shift the position of the COMM domain and make it widely opened to allow the entrance of L-Ser, (4) after the reaction of L-Ser and PLP, the product, E(A-A) intermediate, forms strong HBs with the COMM domain to close the conformation. In Movie S1, we have summarized our findings on the relationship between α and β-ligand binding and the COMM domain opening/closure.
The present study dealt with TRPS from Salmonella typhimurium (St TRPS) because there are many fundamental structural data and compiled knowledge from previous experimental studies. Unfortunately, the sequence similarity between St TRPS and M.tb TRPS is not high, 27% and 57% for the α and β-subunits, respectively. Therefore, the scenario may not be applicable directly to M.tb TRPS for TB drug design. Nonetheless, a similar computation is readily applicable to M.tb TRPS to identify the key site of intersubunit communication.
Although the present study focused on the conformational heterogeneity of TRPS, the atomic details on allosteric chemical reactions, namely, the enhancement of E(Ain).L-Ser →E(A-A) induced by IGP binding and the activation of the α-reaction and indole transfer by the formation of E(A-A), remain an open question. To study the enzymatic reaction, classical molecular dynamics simulation is not enough. We need to perform hybrid QM/MM calculations of both α and β-reactions. (70) Transfer of indole from the α to β-active sites in a narrow tunnel between them provides another computational challenge for us. Without such computational effort, we consider that we are not able to fully understand molecular functions of TRPS and its allosteric regulations. Fortunately, efficient computational methods and sufficient resources are now available to answer these challenging questions.

Supporting Information

ARTICLE SECTIONS
Jump To

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpcb.2c01556.

  • The relationship between α and β-ligand binding and the COMM domain opening/closure (MP4)

    Details on the method and calculations (Tables S1–S3) and additional data (Tables S4–S6, Figures S1–S3) (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

ARTICLE SECTIONS
Jump To

  • Corresponding Author
    • Yuji Sugita - Theoretical Molecular Science Laboratory, RIKEN Cluster for Pioneering Research, 2-1 Hirosawa, Wako, Saitama 351-0198, JapanComputational Biophysics Research Team, RIKEN Center for Computational Science, 7-1-26 Minatojima-Minamimachi, Chuo-ku, Kobe, Hyogo 650-0047, JapanLaboratory for Biomolecular Function Simulation, RIKEN Center for Biosystems Dynamics Research, 1-6-5 Minatojima-Minamimachi, Chuo-ku, Kobe, Hyogo 650-0047, JapanOrcidhttps://orcid.org/0000-0001-9738-9216 Email: [email protected]
  • Authors
    • Shingo Ito - Theoretical Molecular Science Laboratory, RIKEN Cluster for Pioneering Research, 2-1 Hirosawa, Wako, Saitama 351-0198, Japan
    • Kiyoshi Yagi - Theoretical Molecular Science Laboratory, RIKEN Cluster for Pioneering Research, 2-1 Hirosawa, Wako, Saitama 351-0198, JapanOrcidhttps://orcid.org/0000-0003-1120-9355
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

ARTICLE SECTIONS
Jump To

This research is partially supported by RIKEN Pioneering Research Projects (Dynamic Structural Biology/Glycolipidologue Initiative) (to Y.S.), Program for Promoting Research on the Supercomputer Fugaku (Biomolecular dynamics in a living cell (JPMXP 1020200101)/MD-driven Precision Medicine (JPMXP 1020200201)), MEXT/KAKENHI Grants No. JP19H05645, JP21H05249 (to Y.S.) and JP20H02701 (to K.Y.). We used a computer system HOKUSAI (project ID: Q21535), provided by the RIKEN Information System Division, and Oakridge-CX and Octopus, provided by the University of Tokyo and Osaka University, respectively, (hp200098) and Fugaku supercomputer provided by RIKEN (hp210101) through the HPCI System Research Project.

References

ARTICLE SECTIONS
Jump To

This article references 70 other publications.

  1. 1
    Consalvi, S.; Scarpecci, C.; Biava, M.; Poce, G. Mycobacterial tryptophan biosynthesis: A promising target for tuberculosis drug development?. Bioorg. Med. Chem. Lett. 2019, 29, 126731,  DOI: 10.1016/j.bmcl.2019.126731
  2. 2
    Lott, J. S. The tryptophan biosynthetic pathway is essential for Mycobacterium tuberculosis to cause disease. Biochem. Soc. Trans. 2020, 48, 20292037,  DOI: 10.1042/BST20200194
  3. 3
    Global tuberculosis report 2021; World Health Organization, 2021.
  4. 4
    Dunn, M. F. Allosteric regulation of substrate channeling and catalysis in the tryptophan synthase bienzyme complex. Arch. Biochem. Biophys. 2012, 519, 154166,  DOI: 10.1016/j.abb.2012.01.016
  5. 5
    Barends, T. R. M.; Dunn, M. F.; Schlichting, I. Tryptophan synthase, an allosteric molecular factory. Curr. Opin. Chem. Biol. 2008, 12, 593600,  DOI: 10.1016/j.cbpa.2008.07.011
  6. 6
    Wellington, S.; Nag, P. P.; Michalska, K.; Johnston, S. E.; Jedrzejczak, R. P.; Kaushik, V. K.; Clatworthy, A. E.; Siddiqi, N.; McCarren, P.; Bajrami, B. A small-molecule allosteric inhibitor of Mycobacterium tuberculosis tryptophan synthase. Nat. Chem. Biol. 2017, 13, 943950,  DOI: 10.1038/nchembio.2420
  7. 7
    Michalska, K.; Chang, C.; Maltseva, N. I.; Jedrzejczak, R.; Robertson, G. T.; Gusovsky, F.; McCarren, P.; Schreiber, S. L.; Nag, P. P.; Joachimiak, A. Allosteric inhibitors of Mycobacterium tuberculosis tryptophan synthase. Protein Sci. 2020, 29, 779788,  DOI: 10.1002/pro.3825
  8. 8
    Yutani, K.; Ogasahara, K.; Tsujita, T.; Kanemoto, K.; Matsumoto, M.; Tanaka, S.; Miyashita, T.; Matsushiro, A.; Sugino, Y.; Miles, E. W. Tryptophan synthase alpha subunit glutamic acid 49 is essential for activity. Studies with 19 mutants at position 49. J. Biol. Chem. 1987, 262, 1342913433,  DOI: 10.1016/S0021-9258(19)76444-8
  9. 9
    Hyde, C. C.; Ahmed, S. A.; Padlan, E. A.; Miles, E. W.; Davies, D. R. Three-dimensional structure of the tryptophan synthase alpha 2 beta 2 multienzyme complex from Salmonella typhimurium. J. Biol. Chem. 1988, 263, 1785717871,  DOI: 10.1016/S0021-9258(19)77913-7
  10. 10
    Dunn, M. F.; Aguilar, V.; Brzović, P.; Drewe, W. F.; Houben, K. F.; Leja, C. A.; Roy, M. The tryptophan synthase bienzyme complex transfers indole between the α and β sites via a 25-30 Å long tunnel. Biochemistry 1990, 29, 85988607,  DOI: 10.1021/bi00489a015
  11. 11
    Casino, P.; Niks, D.; Ngo, H.; Pan, P.; Brzovic, P.; Blumenstein, L.; Barends, T. R.; Schlichting, I.; Dunn, M. F. Allosteric regulation of tryptophan synthase channeling: The internal aldimine probed by trans-3-indole-3′-acrylate binding. Biochemistry 2007, 46, 77287739,  DOI: 10.1021/bi700386b
  12. 12
    Brzović, P. S.; Ngo, K.; Dunn, M. F. Allosteric interactions coordinate catalytic activity between successive metabolic enzymes in the tryptophan synthase bienzyme complex. Biochemistry 1992, 31, 38313839,  DOI: 10.1021/bi00130a014
  13. 13
    Ngo, H.; Kimmich, N.; Harris, R.; Niks, D.; Blumenstein, L.; Kulik, V.; Barends, T. R.; Schlichting, I.; Dunn, M. F. Allosteric Regulation of Substrate Channeling in Tryptophan Synthase: Modulation of the l-Serine Reaction in Stage I of the β-Reaction by α-Site Ligands. Biochemistry 2007, 46, 77407753,  DOI: 10.1021/bi7003872
  14. 14
    Brzović, P. S.; Sawa, Y.; Hyde, C. C.; Miles, E. W.; Dunn, M. F. Evidence that mutations in a loop region of the α-subunit inhibit the transition from an open to a closed conformation in the tryptophan synthase bienzyme complex. J. Biol. Chem. 1992, 267, 1302813038,  DOI: 10.1016/S0021-9258(18)42377-0
  15. 15
    Brzović, P. S.; Kayastha, A. M.; Miles, E. W.; Dunn, M. F. Substitution of glutamic acid 109 by aspartic acid alters the substrate specificity and catalytic activity of the β-subunit in the tryptophan synthase bienzyme complex from Salmonella typhimurium. Biochemistry 1992, 31, 11801190,  DOI: 10.1021/bi00119a030
  16. 16
    Brzović, P. S.; Hyde, C. C.; Miles, E. W.; Dunn, M. F. Characterization of the functional role of a flexible loop in the α-subunit of tryptophan synthase from Salmonella typhimurium by rapid-scanning, stopped-flow spectroscopy and site-directed mutagenesis. Biochemistry 1993, 32, 1040410413,  DOI: 10.1021/bi00090a016
  17. 17
    Schneider, T. R.; Gerhardt, E.; Lee, M.; Liang, P.-H.; Anderson, K. S.; Schlichting, I. Loop Closure and Intersubunit Communication in Tryptophan Synthase. Biochemistry 1998, 37, 53945406,  DOI: 10.1021/bi9728957
  18. 18
    Jhee, K.-H.; Yang, L.-H.; Ahmed, S. A.; McPhie, P.; Rowlett, R.; Miles, E. W. Mutation of an Active Site Residue of Tryptophan Synthase (β-Serine 377) Alters Cofactor Chemistry. J. Biol. Chem. 1998, 273, 1141711422,  DOI: 10.1074/jbc.273.19.11417
  19. 19
    Ferrari, D.; Yang, L.-H.; Miles, E. W.; Dunn, M. F. βD305A Mutant of Tryptophan Synthase Shows Strongly Perturbed Allosteric Regulation and Substrate Specificity. Biochemistry 2001, 40, 74217432,  DOI: 10.1021/bi002892l
  20. 20
    Kulik, V.; Weyand, M.; Seidel, R.; Niks, D.; Arac, D.; Dunn, M. F.; Schlichting, I. On the Role of αThr183 in the Allosteric Regulation and Catalytic Mechanism of Tryptophan Synthase. J. Mol. Biol. 2002, 324, 677690,  DOI: 10.1016/S0022-2836(02)01109-9
  21. 21
    Kulik, V.; Hartmann, E.; Weyand, M.; Frey, M.; Gierl, A.; Niks, D.; Dunn, M. F.; Schlichting, I. On the Structural Basis of the Catalytic Mechanism and the Regulation of the Alpha Subunit of Tryptophan Synthase from Salmonella typhimurium and BX1 from Maize, Two Evolutionarily Related Enzymes. J. Mol. Biol. 2005, 352, 608620,  DOI: 10.1016/j.jmb.2005.07.014
  22. 22
    Weyand, M.; Schlichting, I. Crystal Structure of Wild-Type Tryptophan Synthase Complexed with the Natural Substrate Indole-3-glycerol Phosphate. Biochemistry 1999, 38, 1646916480,  DOI: 10.1021/bi9920533
  23. 23
    Nishio, K.; Morimoto, Y.; Ishizuka, M.; Ogasahara, K.; Tsukihara, T.; Yutani, K. Conformational Changes in the α-Subunit Coupled to Binding of the β2-Subunit of Tryptophan Synthase from Escherichia coli: Crystal Structure of the Tryptophan Synthase α-Subunit Alone. Biochemistry 2005, 44, 11841192,  DOI: 10.1021/bi047927m
  24. 24
    Ghosh, R. K.; Hilario, E.; Liu, V.; Wang, Y.; Niks, D.; Holmes, J. B.; Sakhrani, V. V.; Mueller, L. J.; Dunn, M. F. Mutation of βGln114 to Ala Alters the Stabilities of Allosteric States in Tryptophan Synthase Catalysis. Biochemistry 2021, 60, 31733186,  DOI: 10.1021/acs.biochem.1c00383
  25. 25
    Marabotti, A.; De Biase, D.; Tramonti, A.; Bettati, S.; Mozzarelli, A. Allosteric Communication of Tryptophan Synthase: Functional and Regulatory Properties of the βS178P mutant. J. Biol. Chem. 2001, 276, 1774717753,  DOI: 10.1074/jbc.M011781200
  26. 26
    Weyand, M.; Schlichting, I.; Herde, P.; Marabotti, A.; Mozzarelli, A. Crystal Structure of the βSer178 → Pro Mutant of Tryptophan Synthase: A “Knock-out” Allosteric Enzyme. J. Biol. Chem. 2002, 277, 1065310660,  DOI: 10.1074/jbc.M111031200
  27. 27
    Lee, T.-S.; Cerutti, D. S.; Mermelstein, D.; Lin, C.; LeGrand, S.; Giese, T. J.; Roitberg, A.; Case, D. A.; Walker, R. C.; York, D. M. GPU-Accelerated Molecular Dynamics and Free Energy Methods in Amber18: Performance Enhancements and New Features. J. Chem. Inf. Model. 2018, 58, 20432050,  DOI: 10.1021/acs.jcim.8b00462
  28. 28
    Phillips, J. C.; Braun, R.; Wang, W.; Gumbart, J.; Tajkhorshid, E.; Villa, E.; Chipot, C.; Skeel, R. D.; Kalé, L.; Schulten, K. Scalable Molecular Dynamics with NAMD. J. Comput. Chem. 2005, 26, 17811802,  DOI: 10.1002/jcc.20289
  29. 29
    Brooks, B. R.; Brooks, C. L.; Mackerell, A. D.; Nilsson, L.; Petrella, R. J.; Roux, B.; Won, Y.; Archontis, G.; Bartels, C.; Boresch, S.; Caflisch, A.; Caves, L.; Cui, Q.; Dinner, A. R.; Feig, M.; Fischer, S.; Gao, J.; Hodoscek, M.; Im, W.; Kuczera, K.; Lazaridis, T.; Ma, J.; Ovchinnikov, V.; Paci, E.; Pastor, R. W.; Post, C. B.; Pu, J. Z.; Schaefer, M.; Tidor, B.; Venable, R. M.; Woodcock, H. L.; Wu, X.; Yang, W.; York, D. M.; Karplus, M. CHARMM: The biomolecular simulation program. J. Comput. Chem. 2009, 30, 15451614,  DOI: 10.1002/jcc.21287
  30. 30
    Shaw, D. E.; Grossman, J. P.; Bank, J. A.; Batson, B.; Butts, J. A.; Chao, J. C.; Deneroff, M. M.; Dror, R. O.; Even, A.; Fenton, C. H.; et al. Anton 2: Raising the Bar for Performance and Programmability in a Special-Purpose Molecular Dynamics Supercomputer. SC ’14: Proceedings of the International Conference for High Performance Computing, Networking, Storage and Analysis, New Orleans, LA, USA, Nov 16–21, 2014; IEEE Press: New Orleans, LA, USA, 2014; pp. 4153.
  31. 31
    Sugita, Y.; Kamiya, M.; Oshima, H.; Re, S. Replica-Exchange Methods for Biomolecular Simulations. Methods Mol. Biol. 2019, 2022, 155177,  DOI: 10.1007/978-1-4939-9608-7_7
  32. 32
    Zuckerman, D. M.; Chong, L. T. Weighted Ensemble Simulation: Review of Methodology, Applications, and Software. Annu. Rev. Biophys. 2017, 46, 4357,  DOI: 10.1146/annurev-biophys-070816-033834
  33. 33
    Valsson, O.; Tiwary, P.; Parrinello, M. Enhancing Important Fluctuations: Rare Events and Metadynamics from a Conceptual Viewpoint. Annu. Rev. Phys. Chem. 2016, 67, 159184,  DOI: 10.1146/annurev-physchem-040215-112229
  34. 34
    Sugita, Y.; Okamoto, Y. Replica-exchange molecular dynamics method for protein folding. Chem. Phys. Lett. 1999, 314, 141151,  DOI: 10.1016/S0009-2614(99)01123-9
  35. 35
    Sugita, Y.; Kitao, A.; Okamoto, Y. Multidimensional replica-exchange method for free-energy calculations. J. Chem. Phys. 2000, 113, 60426051,  DOI: 10.1063/1.1308516
  36. 36
    Sugita, Y.; Okamoto, Y. Replica-exchange multicanonical algorithm and multicanonical replica-exchange method for simulating systems with rough energy landscape. Chem. Phys. Lett. 2000, 329, 261270,  DOI: 10.1016/S0009-2614(00)00999-4
  37. 37
    Kamiya, M.; Sugita, Y. Flexible selection of the solute region in replica exchange with solute tempering: Application to protein-folding simulations. J. Chem. Phys. 2018, 149, 72304,  DOI: 10.1063/1.5016222
  38. 38
    Dokainish, H. M.; Sugita, Y. Exploring Large Domain Motions in Proteins Using Atomistic Molecular Dynamics with Enhanced Conformational Sampling. Int. J. Mol. Sci. 2021, 22, 270,  DOI: 10.3390/ijms22010270
  39. 39
    Oshima, H.; Re, S.; Sugita, Y. Replica-Exchange Umbrella Sampling Combined with Gaussian Accelerated Molecular Dynamics for Free-Energy Calculation of Biomolecules. J. Chem. Theory Comput. 2019, 15, 51995208,  DOI: 10.1021/acs.jctc.9b00761
  40. 40
    Laio, A.; Parrinello, M. Escaping free-energy minima. Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 1256212566,  DOI: 10.1073/pnas.202427399
  41. 41
    Bonomi, M.; Bussi, G.; Camilloni, C.; Tribello, G. A.; Banáč, P.; Barducci, A. Plumed consortium, Promoting transparency and reproducibility in enhanced molecular simulations. Nat. Methods 2019, 16, 670673,  DOI: 10.1038/s41592-019-0506-8
  42. 42
    Yu, I.; Mori, T.; Ando, T.; Harada, R.; Jung, J.; Sugita, Y.; Feig, M. Biomolecular Interactions Modulate Macromolecular Structure and Dynamics in Atomistic Model of a Bacterial Cytoplasm. eLife 2016, 5, 18457,  DOI: 10.7554/eLife.19274
  43. 43
    Jung, J.; Nishima, W.; Daniels, M.; Bascom, G.; Kobayashi, C.; Adedoyin, A.; Wall, M.; Lappala, A.; Phillips, D.; Fischer, W. Scaling molecular dynamics beyond 100,000 processor cores for large-scale biophysical simulations. J. Comput. Chem. 2019, 40, 19191930,  DOI: 10.1002/jcc.25840
  44. 44
    Spyrakis, F.; Raboni, S.; Cozzini, P.; Bettati, S.; Mozzarelli, A. Allosteric communication between α and β subunits of tryptophan synthase: Modelling the open-closed transition of the α subunit. Biochim. Biophys. Acta 2006, 1764, 11021109,  DOI: 10.1016/j.bbapap.2006.03.005
  45. 45
    Fatmi, M. Q.; Ai, R.; Chang, C. A. Synergistic Regulation and Ligand-Induced Conformational Changes of Tryptophan Synthase. Biochemistry 2009, 48, 99219931,  DOI: 10.1021/bi901358j
  46. 46
    Fatmi, M. Q.; Chang, C. A. The Role of Oligomerization and Cooperative Regulation in Protein Function: The Case of Tryptophan Synthase. PLoS Comput. Biol. 2010, 6, e1000994  DOI: 10.1371/journal.pcbi.1000994
  47. 47
    Parveen, T.; Kamran, M.; Fatmi, M. Q. Structural and dynamical thermostability of psychrophilic enzyme at various temperatures: Molecular dynamics simulations of tryptophan synthase. Arch. Biochem. Biophys. 2019, 663, 297305,  DOI: 10.1016/j.abb.2019.01.022
  48. 48
    Maria-Solano, M. A.; Iglesias-Fernández, J.; Osuna, S. Deciphering the Allosterically Driven Conformational Ensemble in Tryptophan Synthase Evolution. J. Am. Chem. Soc. 2019, 141, 1304913056,  DOI: 10.1021/jacs.9b03646
  49. 49
    Zhang, D.; Lazim, R. Exploring indole channeling in tryptophan synthase using steered molecular dynamics simulation. Chem. Phys. Lett. 2019, 734, 136701,  DOI: 10.1016/j.cplett.2019.136701
  50. 50
    Teixeira, C. S. S.; Ramos, M. J.; Sousa, S. F.; Cerqueira, N. M. F. S. A. Solving the Catalytic Mechanism of Tryptophan Synthase: an Emergent Drug Target in the Treatment of Tuberculosis. ChemCatChem 2020, 12, 227237,  DOI: 10.1002/cctc.201901505
  51. 51
    Torrie, G. M.; Valleau, J. P. Nonphysical sampling distributions in Monte Carlo free-energy estimation: Umbrella sampling. J. Comput. Phys. 1977, 23, 187199,  DOI: 10.1016/0021-9991(77)90121-8
  52. 52
    Coutsias, E. A.; Seok, C.; Jacobson, M. P.; Dill, K. A. A kinematic view of loop closure. J. Comput. Chem. 2004, 25, 510528,  DOI: 10.1002/jcc.10416
  53. 53
    Jo, S.; Kim, T.; Iyer, V. G.; Im, W. CHARMM-GUI: A web-based graphical user interface for CHARMM. J. Comput. Chem. 2008, 29, 18591865,  DOI: 10.1002/jcc.20945
  54. 54
    Maier, J. A.; Martinez, C.; Kasavajhala, K.; Wickstrom, L.; Hauser, K. E.; Simmerling, C. ff14SB: Improving the Accuracy of Protein Side Chain and Backbone Parameters from ff99SB. J. Chem. Theory Comput. 2015, 11, 36963713,  DOI: 10.1021/acs.jctc.5b00255
  55. 55
    Wang, J.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. A.; Case, D. A. Development and testing of a general amber force field. J. Comput. Chem. 2004, 25, 11571174,  DOI: 10.1002/jcc.20035
  56. 56
    He, X.; Man, V. H.; Yang, W.; Lee, T.-S.; Wang, J. A fast and high-quality charge model for the next generation general AMBER force field. J. Chem. Phys. 2020, 153, 114502,  DOI: 10.1063/5.0019056
  57. 57
    Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graphics 1996, 14, 3338,  DOI: 10.1016/0263-7855(96)00018-5
  58. 58
    The PyMOL Molecular Graphics System, Version 2.5, Schrödinger: LLC, 2021.
  59. 59
    Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling through velocity rescaling. J. Chem. Phys. 2007, 126, 014101  DOI: 10.1063/1.2408420
  60. 60
    Tuckerman, M.; Berne, B. J.; Martyna, G. J. Reversible multiple time scale molecular dynamics. J. Chem. Phys. 1992, 97, 19902001,  DOI: 10.1063/1.463137
  61. 61
    Darden, T.; York, D.; Pedersen, L. Particle mesh Ewald: AnN·log(N) method for Ewald sums in large systems. J. Chem. Phys. 1993, 98, 1008910092,  DOI: 10.1063/1.464397
  62. 62
    Essmann, U.; Perera, L.; Berkowitz, M. L.; Darden, T.; Lee, H.; Pedersen, L. G. A smooth particle mesh Ewald method. J. Chem. Phys. 1995, 103, 85778593,  DOI: 10.1063/1.470117
  63. 63
    Andersen, H. C. Rattle: A “velocity” version of the shake algorithm for molecular dynamics calculations. J. Comput. Phys. 1983, 52, 2434,  DOI: 10.1016/0021-9991(83)90014-1
  64. 64
    Miyamoto, S.; Kollman, P. A. Settle: An analytical version of the SHAKE and RATTLE algorithm for rigid water models. J. Comput. Chem. 1992, 13, 952962,  DOI: 10.1002/jcc.540130805
  65. 65
    Jung, J.; Kasahara, K.; Kobayashi, C.; Oshima, H.; Mori, T.; Sugita, Y. Optimized Hydrogen Mass Repartitioning Scheme Combined with Accurate Temperature/Pressure Evaluations for Thermodynamic and Kinetic Properties of Biological Systems. J. Chem. Theory Comput. 2021, 17, 53125321,  DOI: 10.1021/acs.jctc.1c00185
  66. 66
    Shirts, M. R.; Chodera, J. D. Statistically optimal analysis of samples from multiple equilibrium states. J. Chem. Phys. 2008, 129, 124105,  DOI: 10.1063/1.2978177
  67. 67
    Chovancova, E.; Pavelka, A.; Benes, P.; Strnad, O.; Brezovsky, J.; Kozlikova, B.; Gora, A.; Sustr, V.; Klvana, M.; Medek, P. CAVER 3.0: A Tool for the Analysis of Transport Pathways in Dynamic Protein Structures. PLoS Comput. Biol. 2012, 8, e1002708,  DOI: 10.1371/journal.pcbi.1002708
  68. 68
    Barends, T. R. M.; Domratcheva, T.; Kulik, V.; Blumenstein, L.; Niks, D.; Dunn, M. F.; Schlichting, I. Structure and Mechanistic Implications of a Tryptophan Synthase Quinonoid Intermediate. ChemBioChem 2008, 9, 10241028,  DOI: 10.1002/cbic.200700703
  69. 69
    Lai, J.; Niks, D.; Wang, Y.; Domratcheva, T.; Barends, T. R. M.; Schwarz, F.; Olsen, R. A.; Elliott, D. W.; Fatmi, M. Q.; Chang, C. A. X-ray and NMR Crystallography in an Enzyme Active Site: The Indoline Quinonoid Intermediate in Tryptophan Synthase. J. Am. Chem. Soc. 2011, 133, 47,  DOI: 10.1021/ja106555c
  70. 70
    Yagi, K.; Ito, S.; Sugita, Y. Exploring the Minimum-Energy Pathways and Free-Energy Profiles of Enzymatic Reactions with QM/MM Calculations. J. Phys. Chem. B 2021, 125, 47014713,  DOI: 10.1021/acs.jpcb.1c01862

Cited By

ARTICLE SECTIONS
Jump To

This article is cited by 3 publications.

  1. Anupom Roy, Mikko Karttunen. A Molecular Dynamics Simulation Study of the Effects of βGln114 Mutation on the Dynamic Behavior of the Catalytic Site of the Tryptophan Synthase. Journal of Chemical Information and Modeling 2024, 64 (3) , 983-1003. https://doi.org/10.1021/acs.jcim.3c01966
  2. Shingo Ito, Kiyoshi Yagi, Yuji Sugita. Allosteric regulation of β-reaction stage I in tryptophan synthase upon the α-ligand binding. The Journal of Chemical Physics 2023, 158 (11) , 115101. https://doi.org/10.1063/5.0134117
  3. Shu Cao, Changbin Tan, Anhua Fei, Gangqiang Hu, Ming Fu, Jun Lv. Insights into pralsetinib resistance to the non-gatekeeper RET kinase G810C mutation through molecular dynamics simulations. Journal of Molecular Modeling 2023, 29 (1) https://doi.org/10.1007/s00894-022-05429-9
  • Abstract

    Figure 1

    Figure 1. (A) Essential domains related to the allostery of TRPS located at the interface of α and β-subunits: αL2 (purple, residue ID 53–60), αL6 (red, residue ID 176–192), αH2 (cyan, residue ID 61–75), and the COMM domain (blue, residue ID 102–189). (B) Entry pathway proposed from the X-ray crystal structures to the β-active site (green). (C) Bottleneck along the proposed L-Ser entry pathway. The white and blue vdW spheres represent the β-subunit and the COMM domain, respectively. The surface of the L-Ser (yellow colored wire) overlaps with the surface of TRPS (vdW sphere). Detailed information on how to set L-Ser in the pathway is described in the Supporting Information.

    Figure 2

    Figure 2. 2D-PMFs along ξCV3 (eq 3) and the RMSD of the COMM domain in reference to an X-ray crystal structure (PDB ID: 2J9X (13)) in the Apo|E(Ain) (A), IGP|E(Ain) (B), and IGP|E(A-A) (C) states. 1D-PMFs along ξCV3 and RMSD are also shown in the bottom and left of each panel, respectively. X-ray crystal structures of Apo|E(Ain) (PDB ID: 1K8X (20)), IGP|E(Ain) (PDB ID: 1WBJ (21)), IGP|E(A-A) (PDB ID: 2J9X, 3CEP, (68) and 3PR2 (69)) states are shown in red, green, and blue circles, respectively. The yellow circle in the IGP|E(Ain) state represents a widely open COMM domain obtained in the present MD simulation.

    Figure 3

    Figure 3. (A) Tunnel from the β-active site to the surface of TRPS (green) of the widely open form in the MD simulations of the IGP|E(Ain) state detected by CAVER 3.0. (67) βD305 and βR141 are shown in red, where the SB is broken in this case. (B) Smallest bottleneck radius was larger than the surface of the L-Ser (yellow wired surface). Detailed information on how to set the L-Ser in the pathway is described in the Supporting Information.

    Figure 4

    Figure 4. Two-dimensional PMFs along ζ (eq 4) and the RMSD of a COMM domain in reference to 2J9X in the Apo|E(Ain) (A), IGP|E(Ain) (B), and IGP|E(A-A) (C) states. 1D-PMFs along ζ and RMSD are also shown in the bottom and left of each panel, respectively. X-ray crystal structures of Apo|E(Ain) (PDB ID: 1K8X), IGP|E(Ain) (PDB ID: 1WBJ), IGP|E(A-A) (PDB ID: 2J9X, 3CEP, and 3PR2) states are shown in red, green, and blue circles, respectively.

    Figure 5

    Figure 5. Probability of HB formation (left) and schematic representations of the HB network (right) in the Apo|E(Ain) and IGP|E(Ain) state. The main HBs with the probability >0.6 are shown in red color. Asterisk indicates the HBs present in the X-ray crystal structures of the IGP|E(Ain) state, 1WBJ.

    Figure 6

    Figure 6. HB network between αL2, αH2, αL6, and the COMM domain. (A–C) HBs between IGP-αD60, αG61-αT183, and αG181-βS178, which are present in the IGP|E(Ain) state (right) and absent in the Apo|E(Ain) state (left). (D, E) HBs between αD60-αQ65, αQ65-βS161, and αP57-βR175 formed in the IGP|E(Ain) state. The structure of the Apo|E(Ain) state is superimposed with a transparent gray color.

    Figure 7

    Figure 7. 1D-PMF of RMSD of COMM domain in the WT–Apo|E(Ain) (top-left), βR175A–Apo|E(Ain) (top right), WT– IGP|E(Ain) (bottom left), and βR175A– IGP|E(Ain) state (bottom right).

    Figure 8

    Figure 8. (A) Probability of HB formation in the IGP|E(Ain) and IGP|E(A-A) states. AA and PP indicate the α-aminoacrylate and phosphate group, respectively. Asterisk represents the HBs present in the X-ray crystal structure, 2J9X. (B, C) HBs between the β-subunit and the β-ligand in the IGP|E(Ain) and IGP|E(A-A) states. (D) Superposition of the conformation in the IGP|E(Ain) state (gray transparent) and IGP|E(A-A) state (colored) around the β-ligand. The two structures were superimposed by all Cα atoms of TRPS except for those of αL2, αL6, αH2, and COMM.

    Figure 9

    Figure 9. RMSD of the COMM domain in the IGP|E(Ain).L-Ser state.

    Figure 10

    Figure 10. Probability of HB formation (A) and the structure (B) in the IGP|E(Ain).L-Ser state.

  • References

    ARTICLE SECTIONS
    Jump To

    This article references 70 other publications.

    1. 1
      Consalvi, S.; Scarpecci, C.; Biava, M.; Poce, G. Mycobacterial tryptophan biosynthesis: A promising target for tuberculosis drug development?. Bioorg. Med. Chem. Lett. 2019, 29, 126731,  DOI: 10.1016/j.bmcl.2019.126731
    2. 2
      Lott, J. S. The tryptophan biosynthetic pathway is essential for Mycobacterium tuberculosis to cause disease. Biochem. Soc. Trans. 2020, 48, 20292037,  DOI: 10.1042/BST20200194
    3. 3
      Global tuberculosis report 2021; World Health Organization, 2021.
    4. 4
      Dunn, M. F. Allosteric regulation of substrate channeling and catalysis in the tryptophan synthase bienzyme complex. Arch. Biochem. Biophys. 2012, 519, 154166,  DOI: 10.1016/j.abb.2012.01.016
    5. 5
      Barends, T. R. M.; Dunn, M. F.; Schlichting, I. Tryptophan synthase, an allosteric molecular factory. Curr. Opin. Chem. Biol. 2008, 12, 593600,  DOI: 10.1016/j.cbpa.2008.07.011
    6. 6
      Wellington, S.; Nag, P. P.; Michalska, K.; Johnston, S. E.; Jedrzejczak, R. P.; Kaushik, V. K.; Clatworthy, A. E.; Siddiqi, N.; McCarren, P.; Bajrami, B. A small-molecule allosteric inhibitor of Mycobacterium tuberculosis tryptophan synthase. Nat. Chem. Biol. 2017, 13, 943950,  DOI: 10.1038/nchembio.2420
    7. 7
      Michalska, K.; Chang, C.; Maltseva, N. I.; Jedrzejczak, R.; Robertson, G. T.; Gusovsky, F.; McCarren, P.; Schreiber, S. L.; Nag, P. P.; Joachimiak, A. Allosteric inhibitors of Mycobacterium tuberculosis tryptophan synthase. Protein Sci. 2020, 29, 779788,  DOI: 10.1002/pro.3825
    8. 8
      Yutani, K.; Ogasahara, K.; Tsujita, T.; Kanemoto, K.; Matsumoto, M.; Tanaka, S.; Miyashita, T.; Matsushiro, A.; Sugino, Y.; Miles, E. W. Tryptophan synthase alpha subunit glutamic acid 49 is essential for activity. Studies with 19 mutants at position 49. J. Biol. Chem. 1987, 262, 1342913433,  DOI: 10.1016/S0021-9258(19)76444-8
    9. 9
      Hyde, C. C.; Ahmed, S. A.; Padlan, E. A.; Miles, E. W.; Davies, D. R. Three-dimensional structure of the tryptophan synthase alpha 2 beta 2 multienzyme complex from Salmonella typhimurium. J. Biol. Chem. 1988, 263, 1785717871,  DOI: 10.1016/S0021-9258(19)77913-7
    10. 10
      Dunn, M. F.; Aguilar, V.; Brzović, P.; Drewe, W. F.; Houben, K. F.; Leja, C. A.; Roy, M. The tryptophan synthase bienzyme complex transfers indole between the α and β sites via a 25-30 Å long tunnel. Biochemistry 1990, 29, 85988607,  DOI: 10.1021/bi00489a015
    11. 11
      Casino, P.; Niks, D.; Ngo, H.; Pan, P.; Brzovic, P.; Blumenstein, L.; Barends, T. R.; Schlichting, I.; Dunn, M. F. Allosteric regulation of tryptophan synthase channeling: The internal aldimine probed by trans-3-indole-3′-acrylate binding. Biochemistry 2007, 46, 77287739,  DOI: 10.1021/bi700386b
    12. 12
      Brzović, P. S.; Ngo, K.; Dunn, M. F. Allosteric interactions coordinate catalytic activity between successive metabolic enzymes in the tryptophan synthase bienzyme complex. Biochemistry 1992, 31, 38313839,  DOI: 10.1021/bi00130a014
    13. 13
      Ngo, H.; Kimmich, N.; Harris, R.; Niks, D.; Blumenstein, L.; Kulik, V.; Barends, T. R.; Schlichting, I.; Dunn, M. F. Allosteric Regulation of Substrate Channeling in Tryptophan Synthase: Modulation of the l-Serine Reaction in Stage I of the β-Reaction by α-Site Ligands. Biochemistry 2007, 46, 77407753,  DOI: 10.1021/bi7003872
    14. 14
      Brzović, P. S.; Sawa, Y.; Hyde, C. C.; Miles, E. W.; Dunn, M. F. Evidence that mutations in a loop region of the α-subunit inhibit the transition from an open to a closed conformation in the tryptophan synthase bienzyme complex. J. Biol. Chem. 1992, 267, 1302813038,  DOI: 10.1016/S0021-9258(18)42377-0
    15. 15
      Brzović, P. S.; Kayastha, A. M.; Miles, E. W.; Dunn, M. F. Substitution of glutamic acid 109 by aspartic acid alters the substrate specificity and catalytic activity of the β-subunit in the tryptophan synthase bienzyme complex from Salmonella typhimurium. Biochemistry 1992, 31, 11801190,  DOI: 10.1021/bi00119a030
    16. 16
      Brzović, P. S.; Hyde, C. C.; Miles, E. W.; Dunn, M. F. Characterization of the functional role of a flexible loop in the α-subunit of tryptophan synthase from Salmonella typhimurium by rapid-scanning, stopped-flow spectroscopy and site-directed mutagenesis. Biochemistry 1993, 32, 1040410413,  DOI: 10.1021/bi00090a016
    17. 17
      Schneider, T. R.; Gerhardt, E.; Lee, M.; Liang, P.-H.; Anderson, K. S.; Schlichting, I. Loop Closure and Intersubunit Communication in Tryptophan Synthase. Biochemistry 1998, 37, 53945406,  DOI: 10.1021/bi9728957
    18. 18
      Jhee, K.-H.; Yang, L.-H.; Ahmed, S. A.; McPhie, P.; Rowlett, R.; Miles, E. W. Mutation of an Active Site Residue of Tryptophan Synthase (β-Serine 377) Alters Cofactor Chemistry. J. Biol. Chem. 1998, 273, 1141711422,  DOI: 10.1074/jbc.273.19.11417
    19. 19
      Ferrari, D.; Yang, L.-H.; Miles, E. W.; Dunn, M. F. βD305A Mutant of Tryptophan Synthase Shows Strongly Perturbed Allosteric Regulation and Substrate Specificity. Biochemistry 2001, 40, 74217432,  DOI: 10.1021/bi002892l
    20. 20
      Kulik, V.; Weyand, M.; Seidel, R.; Niks, D.; Arac, D.; Dunn, M. F.; Schlichting, I. On the Role of αThr183 in the Allosteric Regulation and Catalytic Mechanism of Tryptophan Synthase. J. Mol. Biol. 2002, 324, 677690,  DOI: 10.1016/S0022-2836(02)01109-9
    21. 21
      Kulik, V.; Hartmann, E.; Weyand, M.; Frey, M.; Gierl, A.; Niks, D.; Dunn, M. F.; Schlichting, I. On the Structural Basis of the Catalytic Mechanism and the Regulation of the Alpha Subunit of Tryptophan Synthase from Salmonella typhimurium and BX1 from Maize, Two Evolutionarily Related Enzymes. J. Mol. Biol. 2005, 352, 608620,  DOI: 10.1016/j.jmb.2005.07.014
    22. 22
      Weyand, M.; Schlichting, I. Crystal Structure of Wild-Type Tryptophan Synthase Complexed with the Natural Substrate Indole-3-glycerol Phosphate. Biochemistry 1999, 38, 1646916480,  DOI: 10.1021/bi9920533
    23. 23
      Nishio, K.; Morimoto, Y.; Ishizuka, M.; Ogasahara, K.; Tsukihara, T.; Yutani, K. Conformational Changes in the α-Subunit Coupled to Binding of the β2-Subunit of Tryptophan Synthase from Escherichia coli: Crystal Structure of the Tryptophan Synthase α-Subunit Alone. Biochemistry 2005, 44, 11841192,  DOI: 10.1021/bi047927m
    24. 24
      Ghosh, R. K.; Hilario, E.; Liu, V.; Wang, Y.; Niks, D.; Holmes, J. B.; Sakhrani, V. V.; Mueller, L. J.; Dunn, M. F. Mutation of βGln114 to Ala Alters the Stabilities of Allosteric States in Tryptophan Synthase Catalysis. Biochemistry 2021, 60, 31733186,  DOI: 10.1021/acs.biochem.1c00383
    25. 25
      Marabotti, A.; De Biase, D.; Tramonti, A.; Bettati, S.; Mozzarelli, A. Allosteric Communication of Tryptophan Synthase: Functional and Regulatory Properties of the βS178P mutant. J. Biol. Chem. 2001, 276, 1774717753,  DOI: 10.1074/jbc.M011781200
    26. 26
      Weyand, M.; Schlichting, I.; Herde, P.; Marabotti, A.; Mozzarelli, A. Crystal Structure of the βSer178 → Pro Mutant of Tryptophan Synthase: A “Knock-out” Allosteric Enzyme. J. Biol. Chem. 2002, 277, 1065310660,  DOI: 10.1074/jbc.M111031200
    27. 27
      Lee, T.-S.; Cerutti, D. S.; Mermelstein, D.; Lin, C.; LeGrand, S.; Giese, T. J.; Roitberg, A.; Case, D. A.; Walker, R. C.; York, D. M. GPU-Accelerated Molecular Dynamics and Free Energy Methods in Amber18: Performance Enhancements and New Features. J. Chem. Inf. Model. 2018, 58, 20432050,  DOI: 10.1021/acs.jcim.8b00462
    28. 28
      Phillips, J. C.; Braun, R.; Wang, W.; Gumbart, J.; Tajkhorshid, E.; Villa, E.; Chipot, C.; Skeel, R. D.; Kalé, L.; Schulten, K. Scalable Molecular Dynamics with NAMD. J. Comput. Chem. 2005, 26, 17811802,  DOI: 10.1002/jcc.20289
    29. 29
      Brooks, B. R.; Brooks, C. L.; Mackerell, A. D.; Nilsson, L.; Petrella, R. J.; Roux, B.; Won, Y.; Archontis, G.; Bartels, C.; Boresch, S.; Caflisch, A.; Caves, L.; Cui, Q.; Dinner, A. R.; Feig, M.; Fischer, S.; Gao, J.; Hodoscek, M.; Im, W.; Kuczera, K.; Lazaridis, T.; Ma, J.; Ovchinnikov, V.; Paci, E.; Pastor, R. W.; Post, C. B.; Pu, J. Z.; Schaefer, M.; Tidor, B.; Venable, R. M.; Woodcock, H. L.; Wu, X.; Yang, W.; York, D. M.; Karplus, M. CHARMM: The biomolecular simulation program. J. Comput. Chem. 2009, 30, 15451614,  DOI: 10.1002/jcc.21287
    30. 30
      Shaw, D. E.; Grossman, J. P.; Bank, J. A.; Batson, B.; Butts, J. A.; Chao, J. C.; Deneroff, M. M.; Dror, R. O.; Even, A.; Fenton, C. H.; et al. Anton 2: Raising the Bar for Performance and Programmability in a Special-Purpose Molecular Dynamics Supercomputer. SC ’14: Proceedings of the International Conference for High Performance Computing, Networking, Storage and Analysis, New Orleans, LA, USA, Nov 16–21, 2014; IEEE Press: New Orleans, LA, USA, 2014; pp. 4153.
    31. 31
      Sugita, Y.; Kamiya, M.; Oshima, H.; Re, S. Replica-Exchange Methods for Biomolecular Simulations. Methods Mol. Biol. 2019, 2022, 155177,  DOI: 10.1007/978-1-4939-9608-7_7
    32. 32
      Zuckerman, D. M.; Chong, L. T. Weighted Ensemble Simulation: Review of Methodology, Applications, and Software. Annu. Rev. Biophys. 2017, 46, 4357,  DOI: 10.1146/annurev-biophys-070816-033834
    33. 33
      Valsson, O.; Tiwary, P.; Parrinello, M. Enhancing Important Fluctuations: Rare Events and Metadynamics from a Conceptual Viewpoint. Annu. Rev. Phys. Chem. 2016, 67, 159184,  DOI: 10.1146/annurev-physchem-040215-112229
    34. 34
      Sugita, Y.; Okamoto, Y. Replica-exchange molecular dynamics method for protein folding. Chem. Phys. Lett. 1999, 314, 141151,  DOI: 10.1016/S0009-2614(99)01123-9
    35. 35
      Sugita, Y.; Kitao, A.; Okamoto, Y. Multidimensional replica-exchange method for free-energy calculations. J. Chem. Phys. 2000, 113, 60426051,  DOI: 10.1063/1.1308516
    36. 36
      Sugita, Y.; Okamoto, Y. Replica-exchange multicanonical algorithm and multicanonical replica-exchange method for simulating systems with rough energy landscape. Chem. Phys. Lett. 2000, 329, 261270,  DOI: 10.1016/S0009-2614(00)00999-4
    37. 37
      Kamiya, M.; Sugita, Y. Flexible selection of the solute region in replica exchange with solute tempering: Application to protein-folding simulations. J. Chem. Phys. 2018, 149, 72304,  DOI: 10.1063/1.5016222
    38. 38
      Dokainish, H. M.; Sugita, Y. Exploring Large Domain Motions in Proteins Using Atomistic Molecular Dynamics with Enhanced Conformational Sampling. Int. J. Mol. Sci. 2021, 22, 270,  DOI: 10.3390/ijms22010270
    39. 39
      Oshima, H.; Re, S.; Sugita, Y. Replica-Exchange Umbrella Sampling Combined with Gaussian Accelerated Molecular Dynamics for Free-Energy Calculation of Biomolecules. J. Chem. Theory Comput. 2019, 15, 51995208,  DOI: 10.1021/acs.jctc.9b00761
    40. 40
      Laio, A.; Parrinello, M. Escaping free-energy minima. Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 1256212566,  DOI: 10.1073/pnas.202427399
    41. 41
      Bonomi, M.; Bussi, G.; Camilloni, C.; Tribello, G. A.; Banáč, P.; Barducci, A. Plumed consortium, Promoting transparency and reproducibility in enhanced molecular simulations. Nat. Methods 2019, 16, 670673,  DOI: 10.1038/s41592-019-0506-8
    42. 42
      Yu, I.; Mori, T.; Ando, T.; Harada, R.; Jung, J.; Sugita, Y.; Feig, M. Biomolecular Interactions Modulate Macromolecular Structure and Dynamics in Atomistic Model of a Bacterial Cytoplasm. eLife 2016, 5, 18457,  DOI: 10.7554/eLife.19274
    43. 43
      Jung, J.; Nishima, W.; Daniels, M.; Bascom, G.; Kobayashi, C.; Adedoyin, A.; Wall, M.; Lappala, A.; Phillips, D.; Fischer, W. Scaling molecular dynamics beyond 100,000 processor cores for large-scale biophysical simulations. J. Comput. Chem. 2019, 40, 19191930,  DOI: 10.1002/jcc.25840
    44. 44
      Spyrakis, F.; Raboni, S.; Cozzini, P.; Bettati, S.; Mozzarelli, A. Allosteric communication between α and β subunits of tryptophan synthase: Modelling the open-closed transition of the α subunit. Biochim. Biophys. Acta 2006, 1764, 11021109,  DOI: 10.1016/j.bbapap.2006.03.005
    45. 45
      Fatmi, M. Q.; Ai, R.; Chang, C. A. Synergistic Regulation and Ligand-Induced Conformational Changes of Tryptophan Synthase. Biochemistry 2009, 48, 99219931,  DOI: 10.1021/bi901358j
    46. 46
      Fatmi, M. Q.; Chang, C. A. The Role of Oligomerization and Cooperative Regulation in Protein Function: The Case of Tryptophan Synthase. PLoS Comput. Biol. 2010, 6, e1000994  DOI: 10.1371/journal.pcbi.1000994
    47. 47
      Parveen, T.; Kamran, M.; Fatmi, M. Q. Structural and dynamical thermostability of psychrophilic enzyme at various temperatures: Molecular dynamics simulations of tryptophan synthase. Arch. Biochem. Biophys. 2019, 663, 297305,  DOI: 10.1016/j.abb.2019.01.022
    48. 48
      Maria-Solano, M. A.; Iglesias-Fernández, J.; Osuna, S. Deciphering the Allosterically Driven Conformational Ensemble in Tryptophan Synthase Evolution. J. Am. Chem. Soc. 2019, 141, 1304913056,  DOI: 10.1021/jacs.9b03646
    49. 49
      Zhang, D.; Lazim, R. Exploring indole channeling in tryptophan synthase using steered molecular dynamics simulation. Chem. Phys. Lett. 2019, 734, 136701,  DOI: 10.1016/j.cplett.2019.136701
    50. 50
      Teixeira, C. S. S.; Ramos, M. J.; Sousa, S. F.; Cerqueira, N. M. F. S. A. Solving the Catalytic Mechanism of Tryptophan Synthase: an Emergent Drug Target in the Treatment of Tuberculosis. ChemCatChem 2020, 12, 227237,  DOI: 10.1002/cctc.201901505
    51. 51
      Torrie, G. M.; Valleau, J. P. Nonphysical sampling distributions in Monte Carlo free-energy estimation: Umbrella sampling. J. Comput. Phys. 1977, 23, 187199,  DOI: 10.1016/0021-9991(77)90121-8
    52. 52
      Coutsias, E. A.; Seok, C.; Jacobson, M. P.; Dill, K. A. A kinematic view of loop closure. J. Comput. Chem. 2004, 25, 510528,  DOI: 10.1002/jcc.10416
    53. 53
      Jo, S.; Kim, T.; Iyer, V. G.; Im, W. CHARMM-GUI: A web-based graphical user interface for CHARMM. J. Comput. Chem. 2008, 29, 18591865,  DOI: 10.1002/jcc.20945
    54. 54
      Maier, J. A.; Martinez, C.; Kasavajhala, K.; Wickstrom, L.; Hauser, K. E.; Simmerling, C. ff14SB: Improving the Accuracy of Protein Side Chain and Backbone Parameters from ff99SB. J. Chem. Theory Comput. 2015, 11, 36963713,  DOI: 10.1021/acs.jctc.5b00255
    55. 55
      Wang, J.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. A.; Case, D. A. Development and testing of a general amber force field. J. Comput. Chem. 2004, 25, 11571174,  DOI: 10.1002/jcc.20035
    56. 56
      He, X.; Man, V. H.; Yang, W.; Lee, T.-S.; Wang, J. A fast and high-quality charge model for the next generation general AMBER force field. J. Chem. Phys. 2020, 153, 114502,  DOI: 10.1063/5.0019056
    57. 57
      Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graphics 1996, 14, 3338,  DOI: 10.1016/0263-7855(96)00018-5
    58. 58
      The PyMOL Molecular Graphics System, Version 2.5, Schrödinger: LLC, 2021.
    59. 59
      Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling through velocity rescaling. J. Chem. Phys. 2007, 126, 014101  DOI: 10.1063/1.2408420
    60. 60
      Tuckerman, M.; Berne, B. J.; Martyna, G. J. Reversible multiple time scale molecular dynamics. J. Chem. Phys. 1992, 97, 19902001,  DOI: 10.1063/1.463137
    61. 61
      Darden, T.; York, D.; Pedersen, L. Particle mesh Ewald: AnN·log(N) method for Ewald sums in large systems. J. Chem. Phys. 1993, 98, 1008910092,  DOI: 10.1063/1.464397
    62. 62
      Essmann, U.; Perera, L.; Berkowitz, M. L.; Darden, T.; Lee, H.; Pedersen, L. G. A smooth particle mesh Ewald method. J. Chem. Phys. 1995, 103, 85778593,  DOI: 10.1063/1.470117
    63. 63
      Andersen, H. C. Rattle: A “velocity” version of the shake algorithm for molecular dynamics calculations. J. Comput. Phys. 1983, 52, 2434,  DOI: 10.1016/0021-9991(83)90014-1
    64. 64
      Miyamoto, S.; Kollman, P. A. Settle: An analytical version of the SHAKE and RATTLE algorithm for rigid water models. J. Comput. Chem. 1992, 13, 952962,  DOI: 10.1002/jcc.540130805
    65. 65
      Jung, J.; Kasahara, K.; Kobayashi, C.; Oshima, H.; Mori, T.; Sugita, Y. Optimized Hydrogen Mass Repartitioning Scheme Combined with Accurate Temperature/Pressure Evaluations for Thermodynamic and Kinetic Properties of Biological Systems. J. Chem. Theory Comput. 2021, 17, 53125321,  DOI: 10.1021/acs.jctc.1c00185
    66. 66
      Shirts, M. R.; Chodera, J. D. Statistically optimal analysis of samples from multiple equilibrium states. J. Chem. Phys. 2008, 129, 124105,  DOI: 10.1063/1.2978177
    67. 67
      Chovancova, E.; Pavelka, A.; Benes, P.; Strnad, O.; Brezovsky, J.; Kozlikova, B.; Gora, A.; Sustr, V.; Klvana, M.; Medek, P. CAVER 3.0: A Tool for the Analysis of Transport Pathways in Dynamic Protein Structures. PLoS Comput. Biol. 2012, 8, e1002708,  DOI: 10.1371/journal.pcbi.1002708
    68. 68
      Barends, T. R. M.; Domratcheva, T.; Kulik, V.; Blumenstein, L.; Niks, D.; Dunn, M. F.; Schlichting, I. Structure and Mechanistic Implications of a Tryptophan Synthase Quinonoid Intermediate. ChemBioChem 2008, 9, 10241028,  DOI: 10.1002/cbic.200700703
    69. 69
      Lai, J.; Niks, D.; Wang, Y.; Domratcheva, T.; Barends, T. R. M.; Schwarz, F.; Olsen, R. A.; Elliott, D. W.; Fatmi, M. Q.; Chang, C. A. X-ray and NMR Crystallography in an Enzyme Active Site: The Indoline Quinonoid Intermediate in Tryptophan Synthase. J. Am. Chem. Soc. 2011, 133, 47,  DOI: 10.1021/ja106555c
    70. 70
      Yagi, K.; Ito, S.; Sugita, Y. Exploring the Minimum-Energy Pathways and Free-Energy Profiles of Enzymatic Reactions with QM/MM Calculations. J. Phys. Chem. B 2021, 125, 47014713,  DOI: 10.1021/acs.jpcb.1c01862
  • Supporting Information

    Supporting Information

    ARTICLE SECTIONS
    Jump To

    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpcb.2c01556.

    • The relationship between α and β-ligand binding and the COMM domain opening/closure (MP4)

      Details on the method and calculations (Tables S1–S3) and additional data (Tables S4–S6, Figures S1–S3) (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect