Vaishnavi
Tammara
ab,
Ruchika
Angrover
c,
Disha
Sirur
d and
Atanu
Das
*ab
aPhysical and Materials Chemistry Division, CSIR-National Chemical Laboratory, Dr Homi Bhabha Road, Pune, Maharashtra 411008, India. E-mail: as.das@ncl.res.in
bAcademy of Scientific and Innovative Research (AcSIR), Ghaziabad 201002, India
cThe Departments of the University Institute of Biotechnology, Chandigarh University, NH-05, Ludhiana – Chandigarh State Highway, Punjab 140413, India
dSchool of Physical Sciences, National Institute of Science Education & Research-Bhubaneswar, An OCC of Homi Bhabha National Institute, P.O. Jatni, Khurda, Odisha 752050, India
First published on 30th November 2023
The deleterious impact of Helicobacter pylori (H. pylori) on human health is contingent upon its ability to create and sustain colony structure, which in turn is dictated by the effective performance of flagella – a multi-protein rotary nanodevice. Hence, to design an effective therapeutic strategy against H. pylori, we here conducted a systematic search for an effective druggable site by focusing on the structure–dynamics–energetics–stability landscape of the junction points of three 1:
1 protein complexes (FliFC–FliGN, FliGM–FliMM, and FliYC–FliNC) that contribute mainly to the rotary motion of the flagella via the transformation of information along the junctions over a wide range of pH values operative in the stomach (from neutral to acidic). We applied a gamut of physiologically relevant perturbations in the form of thermal scanning and mechanical force to sample the entire quasi – and non-equilibrium conformational spaces available for the protein complexes under neutral and acidic pH conditions. Our perturbation-induced magnification of conformational distortion approach identified pH-independent protein sequence-specific evolution of precise thermally labile segments, which dictate the specific thermal unfolding mechanism of each complex and this complex-specific pH-independent structural disruption notion remains consistent under mechanical stress as well. Complementing the above observations with the relative rank-ordering of estimated equilibrium binding free energies between two protein sequences of a specific complex quantifies the extent of structure-stability modulation due to pH alteration, rationalizes the exceptional stability of H. pylori under acidic pH conditions, and identifies the pH-independent complex–sequence–segment–residue diagram for targeted drug design.
The MS-ring/C-ring rotor is mainly made up of four proteins – FliG, FliF, FliM, and FliN, and all of them are individually important for the effective formation of the motor5 due to their sequence-specific roles.9–16 Even though there is a consensus on the conserved nature of the structural core of the motor, the similarity in the relative organization of the switch proteins, and the fundamental function of the motor organelle flagella, the diversity among a wide range of microbes originates due to the presence of an additional less abundant protein sequence (FliY).17–20 One such important bacterial species is Helicobacter pylori (H.pylori).19,21 Interestingly, the presence of an additional protein does not alter the core structure of the motor18 and all of them are necessary for the effective performance of the flagella.19,21 Each of these five proteins can be structurally sub-divided into three parts – the N-terminal, middle, and C-terminal domains (denoted with subscripts N, M, and C, respectively). The interactions between a specific pair of proteins take place in a 1
:
1 stoichiometric ratio, e.g., FliFC interacts with FliGN and acts as an anchor for the C-ring;22–24 FliGM is extremely crucial for the interaction with FliMM;25 and FliYC interacts with both FliMC and FliNC.26
We focused on the junction points of the proteins of the MS and C rings of H. pylori (Fig. 1) because – (1) H. pylori is a human pathogenic gut-colonizing bacterium that causes diseases ranging from gastric and peptic ulcer to gastric cancer,27–29 (2) the unique variability in the composition of the motor switch complex of H. pylori is due to the presence of an otherwise less abundant additional protein sequence,26 and (3) the structural integrity of the junctions practically dictates the effective and smooth propagation of information through the complex multi-protein network for precise motor action.25,26,30 Additionally, H. pylori generally exists at neutral or near neutral pH in the stomach and since the stomach can produce very low pH conditions, the change in conditions to acidic pH has the potential to shift the conformational landscape and stability pattern of the overall system and consequently the individual protein complexes and this logical intuition drove us to compare the evolution of conformational landscapes of the complexes under both pH conditions.31 We used a temperature range with high resolution for thermal scanning-mimic simulations32 to understand the structural, dynamical, and energetic features of the complexes, as it helps us to amplify the thermal malleability of a segment and identify the structurally labile fragments that would otherwise have been hidden or sparsely visible at physiological temperature. Going beyond the limit of thermal fluctuations, we further invested our efforts toward the investigation of the maximum potential of inter-protein interactions by applying AFM-mimic external mechanical force.33 Importantly, to draw a conclusion on the dominance and extent of inter-protein interactions in a specific complex, we bridged the gap and hence substantiated our quasi-equilibrium (thermal) and non-equilibrium (mechanical) observations with the thermodynamic parameters in terms of binding affinities between a pair of constituting chains of a particular complex by employing enhanced sampling techniques.34–37 Overall, we cover the entire landscape of the structure-dynamics-energetics-stability-function relationship of the three important complexes that dictate the performance network of the flagellar motor of H. pylori, in turn guiding us to arrive at a specific druggable target site for designing an effective therapeutic strategy against H. pylori, which would be effective in the entire pH range operative in the stomach.
Execution of the above procedure for all three systems gave us a total of thirty systems. Consequently, each one of the thirty systems was put into a rectangular box maintaining a minimum distance of 10 Å between any heavy atom of the system and any edge of the box, which eventually gave us three different box dimensions for the three complexes under investigation – 54.8 × 61.0 × 71.2 Å3 for the FliFC–FliGN complex system, 71.0 × 57.0 × 84.3 Å3 for the FliGM–FliMM complex system, and 66.0 × 59.1 × 67.9 Å3 for the FliYC–FliNC complex system. Then, each system was solvated using the TIP3P water model and electroneutralized using the genion module of GROMACS, where water molecules were replaced randomly by Na+ and Cl− ions such that the salt concentration was maintained at 100 mM. Each of the systems was then energy minimized to reduce any existing steric clashes utilizing the steepest descent algorithm. Furthermore, for every system, two sets of equilibrations were performed consecutively where the center of mass of the system was kept fixed at the center of the box by applying a force constant of 1000 kJ mol−1 nm−2: 500 ps of NVT equilibration at a constant temperature (mentioned above) employing the modified Berendsen thermostat47 and 500 ps of NPT equilibration at a constant temperature (stated above) and at a constant pressure of 1 bar using the modified Berendsen thermostat and Parrinello–Rahman barostat.48 Lastly, a production simulation run for 1 μs was performed for each system removing all position restraints at a specific temperature (discussed above) and at 1 bar pressure. Following are the specifications that were implemented during the production run: an integration time step of 2 fs, 20 ps resolution for snapshot storage, 10 step frequency for the nonbonded interactions list update, the LINCS algorithm49 having a 30° warning angle to restrain bonds with hydrogen atoms, a relaxation constant of 0.1 ps for temperature coupling, a relaxation constant of 0.1 ps and an isothermal compressibility of 4.5 × 10−5 bar−1 for pressure coupling, periodic boundary conditions (PBC) with the standard minimum image convention in all three directions, the Particle Mesh Ewald (PME) method50 to treat long-range electrostatic interactions with a 10 Å real-space cut-off, and a 10 Å cut-off to consider Lennard-Jones interactions with a Fourier spacing of 1.6 Å. The above protocol produces a total of 30 μs of quasi-equilibrium trajectories of the three systems under ten different solution conditions.
Moving on from the global structural to the energetic features, the impact of higher temperature and lower pH becomes extremely consistent irrespective of the species in terms of intra-chain/complex cohesive forces exerted by backbone hydrogen bonds (NHBBB, panel D of Fig. 2 and Fig. S1, S2; panels G and H of Fig. S3–S5, ESI†). Constant-pH and rising temperature or constant-temperature and increasing pH lowers the number of backbone hydrogen bonds, only with a subtle contradicting trend observed for the chains of the FliYC–FliNC complex (Fig. S5G and H, ESI†) at acidic pH in a high-temperature range. However, the impact of acidic pH on creating a stronger network remains sustained for both chains at a specific temperature. Interestingly, intra-complex electrostatic interactions (panel E of Fig. 2 and Fig. S1, S2, ESI†) diminish as pH decreases, and the impact of temperature switches from stabilizing to destabilizing as the pH changes from neutral to acidic. The individual chains (panels I and J of Fig. S3–S5, ESI†) experience a more consistent impact of temperature and pH, as lowering either of them destabilizes the interactions. The above observations are less prominent though for the FliFC–FliGN complex (Fig. 2E) as well as their constituting chains (Fig. S3I and J, ESI†). Similarly, a rise in temperature and an increment in pH generally reduce the van der Waal's stabilization both for the intra-complex (panel F Fig. 2 and Fig. S1, S2, ESI†) and intra-chain scenario (panels K and L of Fig. S3–S5, ESI†); however, a minute inconsistency appears for the FliGM–FliMM complex (Fig. S1F, ESI†) as well as for its constituting chains (Fig. S4K and L, ESI†) at acidic pH and specifically in the lower temperature range.
The intrinsic secondary structural elements of the overall system as well as the individual chains, which were estimated by calculating the probability of specific ϕ–φ dihedral angle combinations evolved during the quasi-equilibrium simulations, that would fall in the α-helix and β-sheet regions according to the Ramachandran plot display a wide range of variations – (1) for the FliFC–FliGN complex (Fig. 2G) and its chains (Fig. S3M and N, ESI†), the secondary structural content decreases with increasing temperature, but remains oblivious to the change in pH, though with a tiny exception for the FliGN chain (Fig. S3N, ESI†) at the highest temperature; (2) for the FliGM–FliMM complex (Fig. S1G, ESI†) and its FliMM chain (Fig. S4M, ESI†), the secondary structural content shows a somewhat consistent decrement at neutral pH with increasing temperature, but a switching point of the trend appears at the acidic pH (prominently for the FliMM chain (Fig. S4M, ESI†)), and generally, acidic pH increases the overall secondary structural content (observed for the FliGM chain (Fig. S4N, ESI†) as well). On the contrary, the FliGM chain (Fig. S4N, ESI†) consistently displays a rise in the secondary structural content with an increment in temperature under both the pH conditions; (3) the FliYC–FliNC complex (Fig. S2G, ESI†) and its FliNC chain (Fig. S5M, ESI†) counterintuitively show an increment in secondary structural content with the rise in temperature and lowering of pH and the FliYC chain (Fig. S5N, ESI†) behaves in a complete reverse fashion.
Further transitioning from the intra-chain to inter-chain parameters, a few trends evolve from the quasi-equilibrium trajectories of all the three complexes – (1) increasing temperature and decreasing pH generally lead to more separation between two chains (measured by calculating the distance between the centers of masses of two interacting chains) of the three complexes (panel H of Fig. 2 and Fig. S1, S2, ESI†), (2) the electrostatic forces operative between a couple of chains always decrease at the acidic pH compared to the neutral one (panel O of Fig. S3–S5, ESI†), but the trend is not so consistent in terms of the impact of temperature, and (3) the lowering of pH always reduces the magnitude of the inter-chain dispersion interactions (panel P of Fig. S3–S5, ESI†) as well, but as observed previously, it displays a pretty inconsistent trend as a function of temperature.
For the FliFC–FliGN complex (Fig. 3A, B and Fig. S6A and B, ESI†), we arrive at the following general mechanism: the initiation of packing alteration leading to swollen ensembles is caused by the rotation of the C-terminus of the FliGN chain (α6-α7); this unpacking event gradually unpacks the N-terminus of the FliGN chain (α1); a combination of the above two events leads to an unfolding of α1 and α7 of the FliGN chain; consequently, the middle segment of the FliGN chain starts unpacking while keeping the FliFC chain almost intact; aggravated unpacking of the mid-part of the FliGN chain lowers the inter-chain interactions, which allows the FliFC chain to distance itself slightly from the FliGN chain; and finally, the chains separate causing substantial unfolding of each of the chains.
For the FliGM–FliMM complex (Fig. 3C, D and Fig. S6C and D, ESI†), in general, the variation initially originates due to the change in the orientation of the first half of α1 of the FliMM chain; consequently, the change in the orientation of α1 propagates to the entire helix, which breaks it into two parts from the middle with a concomitant change in the packing of α2 of the FliMM chain; and finally, the complete loss of β2 and β4 and a partial loss of β3 of the FliMM chain materialize. The afore-mentioned process is non-monotonous as in a few instances few of the secondary structural segments gain percentage, even though partially by a small margin. For the FliGM chain, the diversity originates mainly due to the fluctuating nature of α1.
In the case of the FliYC–FliNC complex (Fig. 3E and F and Fig. S6E and F, ESI†), the first point of differentiation originates due to the out-of-register movement of the C-terminal segment of the α2 of the FliNC chain. The aggravated lack of structural packing between the chains evolves as the mid-part of β1 and N-terminal of β4 of the FliNC chain lose structural integrity. The FliYC chain does not contribute to the initial phase of the perturbations and only displays nominal fluctuations. Consequently, the FliNC chain shows a drastic opening of the N-terminal end of the structure mainly composed of α1, β1, and β2. At this point, the FliYC chain also joins the unfolding phenomenon by showing the complete unfolding of α2 with a concomitant shift of the segment containing β4 and β5 leading to a lack of packing of the conformation, exhibiting aggravated movement of the β1 away from the rest of the protein and indicating a tendency to higher structural shifts for the rest of the secondary structural elements. Finally, both the FliYC and FliNC chains show barely any similarity as further unfolding takes place with residual secondary structural elements.
To further dig deeper into the variability in the evolution of conformational ensembles, we identified the residues belonging to a couple of protein sequences that contribute to the interfacial interactions of a specific complex using PyMOL38 to assess the impact of thermal and pH-driven perturbations (Fig. S7, ESI†). The entire analysis focuses on two major identifiers – the number of residues being involved from each of the protein sequences and the identities of the secondary structural elements that either contribute to or refrain from participating in forming the interface.
In the case of the FliFC–FliGN complex under neutral pH conditions (Fig. S7A, ESI†), the contributions from the two chains in terms of the number of residues (FliFC:
FliGN) change as the temperature increases with the following trend: 28
:
40 (300 K) → 29
:
35 (350 K) → 27
:
40 (400 K) → 24
:
30 (450 K) → 20
:
25 (500 K). The FliFC chain shows more consistency in keeping the interface intact even at elevated temperatures as the contributions come from both α1 and α2 (α2 is the major contributor) and the impact of temperature is more prominent on the interface residues of the FliGN chain. On the contrary, α7 of the FliGN chain never participates in the interface formation (even at 300 K) and as the temperature rises additional secondary structural segments with variable identities pull themselves off from the interface – α1 and α7 at 350 K; α6 at 400 K; α1, α6 and α7 at 450 K; and α1, α5, α6 and α7 at 500 K. The situation is similar under acidic pH conditions (Fig. S7B, ESI†) as well (as expected from the generic unfolding mechanism discussed previously), only with an aggravated scenario. This claim is supplemented by the higher degree of loss of interfacial contacts evidenced by a lower number of residues being involved from each of the chains under acidic pH conditions (FliFC
:
FliGN – 26
:
33 (300 K) → 27
:
35 (350 K) → 23
:
30 (400 K) → 13
:
13 (450 K) → 0
:
0 (500 K)) compared to the neutral pH at a specific temperature. As observed previously for the neutral pH scenario, the FliFC chain participates with both of its α-helices (α2 being the major one) in all four cases, where the interface exists with variable extents (except at 500 K). The FliGN chain shows the following trend in terms of (non)contributing secondary structural elements – non-participating α1 and α7 (300 K); without engaging α1, α6, and α7 (350 K); not involving α1, α4, and α6 (400 K); and contributing via α2 and α5 (450 K) – and the element identities are quite similar to what was observed previously under neutral pH conditions.
Mimicking the same technique for the FliGM–FliMM complex displays the following pattern for the neutral pH condition (Fig. S7C, ESI†) in terms of the number of residues and the secondary structural elements engaging in the interface formation – 23 (α3, α5, α6):
25 (α1, α3, α4) at 300 K, 21 (α3, α5, α6)
:
20 (α1, α3, α4) at 350 K, 17(α1, α2, α6)
:
21 (α1, β1, α4, α5) at 400 K, 9 (α1, α6)
:
19 (α1, β1, β5) at 450 K, and 12 (α2, α5)
:
15 (α1) at 500 K – where the number ratio represents the number of participating residues from the individual chains (FliGM
:
FliMM) and the secondary structural identities are mentioned within the parentheses. It appears that with the rise in temperature, the number of interacting residues decreases due to the increase in the inter-protein distance and the identities of the secondary structural fragments change due to the concomitant structural distortions. The impact of thermal perturbation is even more drastic under acidic pH conditions (Fig. S7D, ESI†) as both fewer residues and hence fewer secondary structural elements participate at any given temperature and the complex dissociates completely from each other from 400 K onwards, which is exemplified by the following quantitative feature – 25 (α3, α4, α5)
:
19 (α4, β4) at 300 K, 8 (α5)
:
12 (β2, β3) at 350 K, and no interaction beyond it.
Since the FliYC–FliNC complex shows the maximum extent of entanglement between two constituting chains, we preferred to explain the interfacial interactions by mentioning the non-participating secondary structural motifs, when necessary, as those would be fewer in numbers. The situation for the complex (FliYC:
FliNC) at the neutral pH (Fig. S7E, ESI†) as a function of rising temperature evolves as follows: at 300 K and 350 K, all the secondary structural elements of both the chains participate in the interface formation with number ratios remaining almost constant – 59
:
58 and 57
:
58 respectively; at 400 K, β3 strands of both chains do not belong to the interface anymore displaying a minute reduction of the interacting residue number ratio (54
:
56); at 450 K, the number of residues is reduced by a substantial amount (47
:
48) with a change in the non-participating secondary structural segment identities (α2 for FliYC and β2 for FliNC); and the situation at 500 K further reduces the number from the previous case (35
:
34), but it reverts to the identity of the non-participating secondary structural elements as observed at 350 K (β3 strands of both chains) with two additional non-engaging secondary structural motifs of the FliNC chain (α1 and β5). The above observation proves that the complex can sustain higher perturbations and keep a formidable number of interfacial interactions and the ability is slightly more prominent for the FliYC chain. Under acidic pH conditions (Fig. S7F, ESI†), interestingly, the events appear as almost similar in terms of identities of participating secondary structural elements only with a higher degree of loss of interfacial interactions at any given temperature – 59 (all)
:
56 (except β3) at 300 K, 55 (all)
:
55 (except β3) at 350 K, 53 (all)
:
48 (except β3) at 400 K, 37 (except α2)
:
36 (all) at 450 K, and 21 (α1, β1, β2, β3)
:
18 (α1 and β1) at 500 K. The change in identities of the participating secondary structural elements can be attributed to the large extent of structural reorganization with progressively applied thermal perturbations.
![]() | ||
Fig. 4 Residue-specific thermal fluctuation profiles evolved under thermal and pH-induced perturbations of the six protein chains. Comparison of the RMSF profiles at a specific pH leads to the identification of thermally fluctuating segments (discussed in the text) that eventually initiate the unfolding of individual chains and consequent separation of the complex-forming chains. In all the panels representing RMSF profiles, the solid lines represent neutral (N) pH conditions and dashed lines represent acidic (A) pH conditions. Moreover, in the case of the RMSF profiles, the five different temperatures used in the study were indicated by the following color scheme – 300 K (blue), 350 K (red), 400 K (yellow), 450 K (purple), and 500 K (green) – irrespective of the pH of the solution. The corresponding structural deviations were shown by superimposing the five most probable states obtained at a specific pH and at five different temperatures for all the three complexes and hence six protein sequences as identified in Fig. 3; the superimposition was performed only in a chain-specific manner (shown as cartoon), rather than in a complex-specific manner, and to have parity with the RMSF profiles in terms of demonstration, the chain colors were chosen according to the temperature from which the coordinates were sampled – (A) and (B) FliFC chain, (C) and (D) FliGN chain, (E) and (F) FliMM chain, (G) and (H) FliGM chain, (I) and (J) FliNC chain, and (K) and (L) FliYC chain. |
The comparison of the RMSF profiles of the FliFC chain of the FliFC–FliGN complex (Fig. 4A) displays heterogeneous increments of residue-wise thermal fluctuations under two different pH conditions. Estimation of the difference between the profiles under a given pH condition leads to the identification of dominant fluctuating residues under two different pH conditions (Fig. 4B). Interestingly, the identities of the dynamics-dictating residues and/or segments overlap between neutral (I537–K542 and P545–E547) and acidic pH (K536–K542 and P545–A549) conditions and suggest that residues close to either side of the loop connecting α1-α2 dominate the dynamical evolution pattern of the FliFC chain. The corresponding DCCM analysis (Fig. S8A–J, ESI†) of the FliFC chain attributes the impact of higher temperatures and lower pH values to the gradual evolution of small, correlated domains throughout the chain compared to the helical-domain-specific correlations and further appearance of strong anti-correlated movements between the two helices, which eventually leads to structural opening. A similar approach applied for the FliGN chain displays the comparison of the RMSF profiles (Fig. 4C) and identifies the dynamically important residues (Fig. 4D). Under neutral pH conditions, two segments emerge as dynamically dominant regions – one with the N-terminal part of α4 (L41–E48) and the other with the entire α5 and the loop connecting α5 and α6 (Q62–N82). Acidic pH does not alter the number of dominating segments but changes the locations, though with some overlap – one with the loop connecting α4 and α5 and the N-terminal part of α5 (N57–V67) and the other with the C-terminal part of α5 and the loop connecting α5 and α6 (I74–G84). The structural unfolding of the FliGN chain appears when the DCCM pattern (Fig. S8K–T, ESI†) changes from the appearance of numerous numbers of small correlated and anti-correlated domains to the genesis of marginal longer ones, specifically when the entire segment α1–α4 is highly correlated with itself and extremely anti-correlated with the rest of the chain – mainly from the mid of α4 to the mid of α6.
Interestingly, the comparison of the RMSF profiles of the FliMM chain of the FliGM–FliMM complex (Fig. 4E) and the corresponding identification of the dynamical protagonists (Fig. 4F) show that the number of dominating segments get reduced on going from the neutral (5 segments) to the acidic pH (2 segments). At the neutral pH, the five segments that contribute mainly to the dynamical evolution of the FliMM conformations are the entire β1 strand (E80–M89), β2 and part of β3 (P101–G116), the loop connecting α3 and α4 and the N-terminal part of α4 (K131–D155), the complete β4 strand (F176–S184), and the C-terminal part of β5 and the N-terminal part of β6 and the loop connecting them (E204–G212). On going from the neutral to the acidic pH, the important segments get focused down to two segments – one with β1, α2, β2, and part of β3 (S77–G116) and the other with the C-terminal part of the α4 and the entire β4 strand (K166–S185). Although it seems like there is a drastic reduction in the number of dominating segments on going from the neutral to the acidic pH, practically this originates due to the merger of segments one and two and the extension of segment four at acidic pH with respect to the neutral one, proving the fact that there is a huge overlap between dominating residues under two different pH conditions. The DCCM profiles (Fig. S9A–J, ESI†) overall state that the unfolding of the FliMM chain takes place when the segment containing α1–α3 and β1–β3 is extremely correlated and shows strong dynamical anti-correlation with the rest of the chain. For the FliGM chain (Fig. 4G and H) however things are much simpler as two common dominating segments appear under two different pH conditions, only with slight variations in terms of length – one stretching from the C-terminal part of α2 to the N-terminal half of α3 (F133–I142) and the other constituting the entire α5 (F161–R172); only at acidic pH, an additional region appears, which stretches from the C-terminal part of α3 to the N-terminal half of α4 (L146–A154), and the previously observed two segments are shortened (I135–T141 and I169–L176). Here, the structural unfolding happens when the region stretching from α1–α3 is highly dynamically correlated with itself and shows complete anti-correlation with the regions stretching from α4–α7 (Fig. S9K–T, ESI†).
In the case of the FliYC–FliNC complex, the corresponding comparisons of the RMSF profiles and labile dynamically dominating segments are shown in Fig. 4I and J for the FliNC chain and Fig. 4K and L for the FliYC chain, respectively. For the FliNC chain, the major contributing segments overlap nicely even if the solution condition changes from neutral to acidic pH. Overall, three major segments appear as main protagonists – a segment stretching the entire length of the β2 strand and the loop connecting α1 and β2 (E66–I71/F65–P76), a sequence containing the β3 strand and the loop connecting β3 and β4 (V85-G91/D82-I90), and a final segment that stretches from the loop connecting the β4 and β5 strands and the N-terminal segment of the β5 strand (F98-R105/V97–A103) – obviously with nominal fluctuations in terms of their lengths under different solution conditions. Similarly, for the FliYC chain, the contributing segments toward partial unfolding mainly overlap under different pH conditions. Although it seems that lowering the pH increases the number of thermally malleable segments from three (D232–N244, I249–E260, and D265–Q271 at neutral pH) to four (M231–V236, L242-P246, E248–A257, and V262–N267 at acidic pH), practically it is just a breakage of the continuity of the first labile segment observed at neutral pH into two fluctuating fragments at acidic pH. Broadly, the malleable parts can be classified in terms of their content of secondary structural elements as below: the loop connecting α1 and β2; the β2 strand; the loop joining β2 and β3; the entire segment containing β3 and β4; and finally, the loop connecting β4 and β5 and the N-terminal end of β5. In terms of dynamical communication, the DCCM analysis of the FliNC chain (Fig. S10A–J, ESI†) shows that the unfolding initiates due to a strongly correlated motion within a continuous stretch of residues consisting of β3, β4, β5, and α1. However, the extensive unfolding appears when the N-terminal segment of the chain containing β1 moves in an anti-correlated fashion with respect to the rest of the chain. The extent of anti-correlation could be higher with the mid-part of the chain containing α1, β2, β3, and β4 or the C-terminal end of the chain consisting of β5 and α2. On the other hand, for the FliYC chain, a similar cross-correlation analysis (Fig. S10K–T, ESI†) shows that the initiation of the unfolding of the chain develops due to a strongly correlated dynamical segment with β2, β3, and β4 strands. In this case, the extensive unfolding can appear only when there is a strong dynamically correlated domain containing α1, which is highly anti-correlated with the rest of the chain – specifically β2 and a segment containing β4, β5, and α2.
Similar to the previously observed trends, acidic pH generally breaks the inter-chain hydrogen bond network (Fig. 5C) at a faster rate compared to neutral pH for any given complex. Intriguingly, the relative rank ordering of the absolute values of the inter-chain hydrogen bonds (FliYC–FliNC > FliFC–FliGN > FliGM–FliMM) rationalizes the previously observed rank ordering of both the rupture force values and the separation initiation time scales, as it is directly correlated to each of those parameters. The extent of operative inter-chain non-bonding interactions, whether electrostatic (Fig. 5D) or van der Waals (Fig. 5E), follows the same pattern as observed previously for all the three reaction coordinates (force, distance, and hydrogen bond), i.e., FliYC–FliNC > FliFC–FliGN > FliGM–FliMM irrespective of the pH conditions. Similarly, the impact of pH remains consistent for both the energetic terms, as acidic pH generally reduces the operative forces – be it electrostatic or van der Waals. However, for a given complex, the lowering of pH reduces the inter-chain electrostatic interaction by a much larger factor and that impact is visible in the case of van der Waals forces, but to a much lesser extent. This observation can be attributed to the excessive positive potential of the interacting chains at acidic pH, leading to the generation of electrostatic repulsion between similar types of charges and finally producing less favorable cohesive inter-chain interactions.
The above set of analyses lead to a conclusion that the combined impact of external mechanical force and acidic pH follows an additivity rule as lowering the pH of the solution generally reduces the cohesive interactions and hence the mechanical stability of each of the three complexes. To further decipher the mechanism of the complexes in terms of structural visualization, we chose intermediate conformations from the pulling assays and depicted them as a function of a progressive time scale. Fig. 5F shows the variability in the evolution of the structural features of the FliFC–FliGN complex under both solution conditions. The general theme that appears in the mechanical separation process of the FliFC–FliGN complex is the retainment of the structural integrity of the FliGN chain irrespective of the pH conditions. On the contrary, the FliFC chain shows a loss of compactness of the native fold along with a concomitant loss in secondary structural features and the differentiation of the separation pathways under two different pH conditions mainly arises due to the conformational variability of the FliFC chain. On the other hand, neither of the two chains of the FliGM–FliMM complex shows any hint of loss of global fold as well as loss of secondary structures during the mechanical unfolding process independent of the pH of the solution (Fig. 5G). It seems that during the separation process, the FliGM and FliMM chains merely glide past each other and this observation is in accordance with the less amount of cohesive interactions operating between the chains as observed previously. Interestingly, for the FliYC–FliNC complex, both chains show a loss of structural integrity in terms of tertiary and secondary structural features irrespective of the pH conditions present (Fig. 5H). Two major important observations evolve in the mechanical separation process of the FliYC–FliNC complex – (1) irrespective of the pH conditions, the loss in structural feature is more prevalent in the FliNC chain and (2) under acidic pH conditions, the FliNC chain shows an even aggravated loss of secondary structural features. The latter might be attributed to the lowered inter-chain interactions observed for the FliYC–FliNC complex under acidic pH conditions.
To add to the categorization of the mechanical unfolding mechanisms, the distance between the constituting chains was computed, however, by deconvoluting them into four possible combinations, where, as a combination, the analysis shows how the two ends of a protein sequence get separated from the other two ends of the complementary complex-forming sequence (Fig. S11, ESI†). For the FliFC–FliGN complex, the separation gets initiated by the N-terminus of the FliFC chain as it starts moving away simultaneously from the N- and C-termini of the FliGN chain and this is further followed by the second event of separation where the C-terminal end of the FliFC chain gets separated from the FliGN chain as well leading to the complete separation of the two proteins (Fig. S11A, ESI†). This suggests that the two proteins of the FliFC–FliGN complex undergo mechanical separation via the sequential unzipping mechanism, i.e., the separation of two interacting chains of a specific complex initiates preferentially at one end of the interface and gradually this structural perturbation permeates to the other end of the interface, and interestingly, the mechanism remains consistent under two different pH conditions. In the case of the FliGM–FliMM complex, however, the separation event does not show any distinctive bias via any end of the separating strands and hence it can be termed as the simultaneous separation mechanism, which remains sustained irrespective of the pH of the solution (Fig. S11B, ESI†). When it comes to the FliYC–FliNC complex, the mechanism slightly changes on going from the neutral to the acidic pH (Fig. S11C, ESI†). At neutral pH, the separation gets initiated by the C-terminus of the FliNC chain from the C-terminus of the FliYC chain, which gets further aggravated by the simultaneous separation of the N- and C-termini of the FliNC chain from the C- and N-termini of the FliYC chain, respectively, and which finally gets completed by the separation of the two N-termini of the FliYC and FliNC chains. Under acidic pH conditions, the initial and final separation events remain similar, and only the middle part of the event where the N- and C-termini of the FliNC chain get separated from the C- and N-termini of the FliYC chain varies from the neutral pH conditions as these two events happen sequentially and not simultaneously as observed previously. Overall, the mechanical separation follows a sequential unzipping mechanism in the FliYC–FliNC complex irrespective of the solution conditions.
Under the thermal and pH-dependent perturbations, the beginning of the structural deformation is exemplified by a subtle gain in backbone compactness with concomitant retainment of the overall solvent exposure of a protein/complex hinting toward the generation of a fluffy state57 where labile side chains are loosely packed to the core of the protein and this phenomenon becomes more prominent for α-helix-rich systems. Post-attainment of the fluffy state, the progressive loss of structural integrity comprises an anticipated route with a higher degree of structural unfolding where all the structural order parameters correlatively tell the same story. On the contrary, the energetic features whose balance helps to hold the structural integrity of a protein sequence display an extremely predictable trend of gradual weakening of energetic stabilization, specifically in terms of the backbone hydrogen bond network58 and dispersion interactions.59 The somewhat counterintuitive trend as observed for the long-range electrostatic interactions under a specific solution condition might be due to the emergence of non-specific interactions within/between protein sequences because of the creation of more exposed conformational ensembles and hence the increased possibility of intra-/inter-chain interactions.60 However, the trend also becomes intuitive when two different solution conditions are compared as acidic pH always destabilizes the system electrostatically – be it intra-protein or inter-protein.61 A similar dual character is displayed by secondary structural content – intuitive (FliFC–FliGN system), as aggravated perturbing conditions always reduce it at a specific pH, and counter-intuitive (FliGM–FliMM and FliYC–FliNC systems), as the earlier trend becomes inconsistent. In the first case, the percentage of secondary structure remains unchanged due to the protonation of side chains (keeping the other conditions unchanged) and this hints that the system loses it tertiary structure at acidic pH, which leads to less packing of the side chains (complementing structural order parameters).62 For the latter scenario, the gain in secondary structure with progressively aggravated perturbations under a particular solution condition or due to the lowering of pH without altering the rest of the solution conditions could be attributed to the preferential unfolding of α-helical segments and creating a tendency to generate β-sheet-like structures,63 at least in terms of the ϕ–φ landscape. Additionally, in general, the characteristics of each of these three multi-protein complexes are mainly dictated by its longer protein fragment. Importantly, irrespective of the nature of perturbations being implemented (elevated temperature or low pH), each of the three complexes exhibits a complex-specific generic unfolding mechanism whose sequence of events remains unaltered proving the existence of a complex-and chain-specific maintenance of structural integrity, and this claim is further substantiated by the observed similarity in the modulation of interfacial interactions of a specific complex under thermal/pH-driven perturbations. The combined effect of high temperature and low pH perturbations reduces the withstanding ability of a complex to a much greater extent due to the additive nature of the combined external force acting on the structure.64
One of the most important features of the present work is the identification of malleable segments that are prone to extensive fluctuations under thermal and pH-modulated perturbations and hence can act as the potential sites for the initiation of structural disintegration.65 Interestingly, the specific segments that appear as labile fragments of a sequence retain their identities in terms of the length and position when solution conditions are changed to the acidic pH, which practically suggests that each protein sequence has multiple generic discrete sites that are vulnerable to any form of cellular perturbations (weak spots) and would start unfolding early. The protein-specific network of dynamical cross-talk percolates these environmental heterogeneity-driven segment-specific perturbations to the entire sequence. The progressive loss of structural integrity of a protein always involves the melting of a large number of smaller dynamical networks to a very few numbers of continuous networks where torque is generated by the out-of-phase dynamical movement of a considerably big part of a chain from the rest of it.66
Application of mechanical perturbation in our systems makes sense as the motor proteins experience conversion of electrochemical potential to mechanical force in the process of generating torque.30 Interestingly, the complexes show variability in their mechanical separation processes by projecting two different mechanisms – sequential unzipping and simultaneous separation67 – and more importantly, a particular complex does not change its mechanical separation mechanism with varying pH. It is important here to mention that the observed mechanical separation mechanisms of the three individually studied complexes might vary if the entire oligomeric motor protein complex is considered simultaneously where individual complexes are connected as there could be some long-range allosteric impact of one complex on another. Since the conformations evolved in the pulling trajectories are non-equilibrium in nature as well as the thermally perturbed ensembles, a direct comparison between them would not make much sense statistically. Interestingly though, the additivity feature of combined perturbations is found to be valid for mechanical perturbation and acidic pH as well.68
The limitation of quasi- to non-equilibrium sampling is overcome by employing enhanced sampling approaches which eventually led us to the equilibrium quantification of binding affinities between two partnering protein sequences of a specific complex. In general, acidic pH lowers the binding affinity between two protein chains of a complex compared to the neutral one and this in turn practically provides robustness to the quasi- and non-equilibrium measurements of structural and energetic variables, which hinted toward the reduction of cohesive forces operative between two partnering proteins of a complex under acidic pH conditions. It is worth emphasizing that the reduction of affinity between a pair of protein chains under acidic pH conditions varies significantly as the affinity gets reduced to a very large extent for the FliGM–FliMM complex, but not so much for the other two complexes. Specifically, the FliFC and FliGC chains, which form the MS-ring/C-ring junction point, get minimally affected due to the lowering of pH and this might be a testimony of the unperturbed functioning of H. pylori under acidic pH conditions in the stomach.31 The extremely lowered affinity between the FliGM and FliMM proteins under acidic pH conditions does not show any considerable impact on the flagellar functional activity31 and this might be due to the nature of the structure of the FliGM–FliMM complex, where the structure-forming proteins are merely held together without any cross-structure entanglements (a larger distance between two partnering proteins), unlike what is observed for the other two complexes (a smaller distance between the structure-forming proteins). The correlation between estimated binding affinities at neutral pH and the extent of entanglement between two partnering protein sequences of a specific complex in their native states is further evidenced by the number of residues involved in the interfacial interactions, which is maximum for the FliYC–FliNC complex and minimum for the FliGM–FliMM complex.
Although there have been extensive research reports on the H. pylori flagellar motor system, a lack of detail is eminent when it comes to the extremely specific structural and energetic features of the constituting proteins or the junction complexes down to the single molecule level and this presents a hurdle for any simulated observable to be compared and hence to be validated. Still, we found multiple instances where our simulated observables qualitatively/semi-quantitatively match with the existing experimental observations: the highest extent of structural deviation of two helices (α6 and α7) of the FliGN chain in the FliFC–FliGN complex,69 which might be due to the fact that those two helices are the loosely bound segments as they form after the complex formation between FliF and FliG;30 acidic pH drives the FliFC–FliGN complex more towards the dissociated state;70 FliYC–FliNC inter-chain interactions are dominated by electrostatic forces;26 both polar and hydrophobic interactions are essential for the FliGM–FliMM complex to be in the stable state;25 protonation leads to reduced inter-chain interactions for the FliFC–FliGN complex due to lack of polar interactions and residual interactions are contributed by the hydrophobic residues;30 key residues/secondary structural elements contributing to the interfacial interactions of three complexes in their native states corroborate nicely with experimentally determined structural information;25,26,30 α5 of FliGN, which is reported to be important for flagellar formation and motility,30 evolves as a malleable segment in our analysis; the loop connecting α5 and α6, which supposedly prevents intramolecular association,30 also appears as a stand out segment in fluctuation profile analysis; cross-species conserved residues25 of both FliGM (E137, H138 and R172) and FliMM (R76, D130, L133, R144, S147 and I149) appear to be labile segments in our study as well; and the relative order of binding affinities between FliFC–FliGN and FliGM–FliMM complexes, as estimated by our simulated approach, matches nicely with the experimentally determined relative trend of dissociation constants of these two complexes.25,30 All the above comparisons add robustness to our inferences drawn from simulated trajectories.
Footnote |
† Electronic supplementary information (ESI) available: Structural and energetic features obtained from the quasi-equilibrium trajectories of two complexes (FliGM-FliMM and FliYC-FliNC) and six chains (FliFC, FliGN, FliGM, FliMM, FliYC, and FliNC), generic unfolding mechanisms of three complexes under thermal and/or pH-driven perturbations, modulation of interfacial interactions of three complexes as a function of temperature and pH, intra-chain DCCMs of six chains (FliFC, FliGN, FliGM, FliMM, FliYC, and FliNC), terminus-specific inter-chain distance for three systems (FliFC-FliGN, FliGM-FliMM and FliYC-FliNC), and robustness check for umbrella sampling simulations for the three systems (FliFC-FliGN, FliGM-FliMM and FliYC-FliNC). See DOI: https://doi.org/10.1039/d3cp05024f |
This journal is © the Owner Societies 2024 |