Skip to main content

This is a preprint.

It has not yet been peer reviewed by a journal.

The National Library of Medicine is running a pilot to include preprints that result from research funded by NIH in PMC and PubMed.

bioRxiv logoLink to bioRxiv
[Preprint]. 2025 Jun 17:2024.07.22.604679. Originally published 2024 Jul 23. [Version 2] doi: 10.1101/2024.07.22.604679

Transient Non-local Interactions Dominate the Dynamics of Measles Virus NTAIL

Lillian Otteson 1,2,&, Gabor Nagy 3,&, John Kunkel 1,2, Gerdenis Kodis 1,2, Lars V Bock 3, Christophe Bignon 5, Sonia Longhi 5, Wenwei Zheng 4,*, Helmut Grubmüller 3,*, Andrea C Vaiana 3,6,*, Sara M Vaiana 1,2,*
PMCID: PMC11291014  PMID: 39091801

Abstract

The RNA genome of measles virus is encapsidated by the nucleoprotein within a helical nucleocapsid that serves as template for both transcription and replication. The intrinsically disordered domain of the nucleoprotein (NTAIL), partly protruding outward from the nucleocapsid, is essential for binding the polymerase complex responsible for viral transcription and replication. As for many IDPs, binding of NTAIL occurs through a short molecular recognition element (MoRE) that folds upon binding, with the majority of NTAIL remaining disordered. Though NTAIL regions far from the MoRE influence the binding affinity, interactions between them and the MoRE have not been investigated in depth. Using an integrated approach, relying on photo-induced electron transfer (PET) experiments between tryptophan and cysteine pairs placed at different positions in the protein under varying salt and pH conditions, combined with simulations and analytical models, we identified transient interactions between two disordered regions distant in sequence, which dominate NTAIL dynamics, and regulate the conformational preferences of both the MoRE and the entire NTAIL domain. Co-evolutionary analysis corroborates our findings and suggests an important functional role for the same intramolecular interactions. We propose mechanisms by which these non-local interactions may regulate binding to the measles phosphoprotein, polymerase recruitment, and ultimately viral transcription and replication. Our findings may be extended to other IDPs, where non-local intra-protein interactions affect the conformational preferences of intermolecular binding sites.

Introduction

The measles virus (MeV) is a non-segmented, single-stranded, negative sense RNA virus, in which the viral genome is wrapped within a helical nucleocapsid, made of thousands of repeats of the nucleoprotein (N)1. This structure is shared among all Paramyxo- and Pneumoviruses, which include common human viruses (e.g., MeV, mumps, respiratory syncytial virus, metapneumovirus, and other para-influenza viruses) and animal viruses (e.g., Sendai virus, Newcastle disease virus and rinderpest virus), as well as zoonotic biosafety level 4 agents of great concern to public health (e.g., Hendra and Nipah viruses, respectively HeV and NiV)2. The N protein not only provides the interaction with the RNA and holds the nucleocapsid in its helical structure, but it is the substrate used for viral transcription and replication. Viral transcription and replication in Paramyxoviruses are ensured by a complex formed by the large-polymerase (L) and the phosphoprotein (P), where the latter enables tethering L to the nucleoprotein-RNA (N-RNA) template for transcription and replication2,3.

In MeV, N consists of a folded 400 amino-acid (a.a.) domain (NCORE) which binds the RNA and forms the structured helical nucleocapsid, and of an intrinsically disordered 125 aa. C-terminal domain (NTAIL), which partly protrudes radially outward from the helical nucleocapsid structure47. The intrinsically disordered NTAIL domain binds to the folded X domain (PXD) of the phosphoprotein (P) and is essential for MeV transcription and replication. P in turn, binds as a tetramer to both L and NCORE6,811. A similar structure and binding of NTAIL to PXD are found in NiV and HeV3,12. Binding of NTAIL to PXD is thought to assist in positioning the L appropriately on the nucleocapsid to read the RNA, while allowing efficient transcription re-initiation at intergenic regions13 and hence progression from one reading frame to the next. It may also assist in maintaining contact between L and the nucleocapsid template, preventing it from leaving before transcribing all the viral genome11,14. In addition, the presence of NTAIL loosens the nucleocapsid helical structure, possibly facilitating transcription and replication by L11.

Because of its crucial role in transcription and replication, binding of MeV NTAIL to PXD has been extensively studied1521. It occurs through coupled folding and binding of a short molecular recognition region (MoRE) of 18 a.a. (grey in Fig. 1)21. Upon binding, the MoRE folds into an α-helix, forming a four-helix bundle with three helices from PXD (PDB: 1t6o8,9). However, the majority of NTAIL (i.e., the remaining 109 a.a. on the two sides of the MoRE) remains disordered in the bound complex21. The NTAIL-PXD complex is structurally similar in MeV, NiV and HeV3,12. While the regions outside the MoRE (from here on referred to as “flanking regions”) do not fold, or engage in stable binding interactions16,17, they modulate binding affinities, as indicated by the binding affinity increase in truncated variants of NTAIL containing the intact MoRE22,23. This increase might be due to a pure excluded volume (entropic) repulsion of the flanking regions, as well as to potential interactions involving the flanking regions of NTAIL (either at the intra- or inter-protein level)22,24,25.

Figure 1.

Figure 1.

Results of photo-induced electron transfer (PET) experiments at T=20 °C, on NTAIL variants containing a unique C-W pair. A) NTAIL sequence, showing charged residues at pH 7.6, location of Y to W, and S to C substitutions, and location of MoRE, which folds into a helix upon binding to PXD. B) Net charge per residue calculated using a window size of six amino acids. C) W and C positions. Each variant contains a unique C-W pair. The C-W distance in each variant is indicated by orange dashed lines. C3W2 and C4W2 variants encompass the MoRE transient helix. D) C-W relaxation time τCW, due to PET as a function of C-W sequence separation. Experimental conditions: pH 7.6 and 150 mM NaCl (black circles); pH 4.0 and 150 mM NaCl (red squares); pH 7.6 and 500mM NaCl (green triangles). Dashed line: expected scaling of τCW, from a SAW-ν homopolymer model at pH 7.6 (black). Inset: Comparison of NTAIL circular dichroism spectra at pH 7.6 (black) and pH 4 (red).

For NTAIL free in solution, the secondary structure propensity of the MoRE region has been characterized by NMR on full length NTAIL7, and by all-atom simulation studies of the MoRE fragment19,26. In contrast, we know very little about potential interactions between residues that are distant in sequence (non-local interactions), despite their particular importance for NTAIL binding to PXD24,27 and, more generally, for IDP function. Large-scale reconfiguration dynamics can regulate IDP binding mechanisms, rates, and affinities; in fact, examples of this are found even in simple models of proteins undergoing two-state conformational transitions28. For NTAIL, we expect interactions involving the regions flanking the MoRE on either side to confer binding specificity to the IDP by dictating large-scale reconfiguration dynamics. In addition, non-local interactions potentially act as allosteric regulators in IDPs by stabilizing certain conformational states. The resulting shift in the population of states within the ensemble can lead to allosteric regulation, as described in the ensemble allosteric model for IDPs29,30.

Probing non-local intra-chain interactions is therefore essential for a molecular understanding of IDP dynamics, binding, and function in general22,3133. Direct experimental probing of such interactions is particularly difficult, due to their transient nature, and the fact that IDPs undergo large conformational changes on sub-microsecond time scales. Atomistic simulations of IDPs, on the other hand, are equally challenging due to the wide conformational space sampled by IDPs. To overcome these experimental and theoretical challenges, we leveraged the high sensitivity of IDP dynamics to non-local interactions. We experimentally quantified the dynamics of full-length NTAIL in solution, by probing intra-molecular contact formation times, using photo-induced electron transfer (PET) between a tryptophan (W) and a cysteine (C) placed in the sequence3437. Using this technique, we measured contact formation rates along the well-defined C-W distance reaction coordinate, probing large-scale backbone motions on the hundreds of nanoseconds to microsecond internal reorganization time scale36, which is most relevant to IDP function38. With minimal perturbation to the sequence, compared to introducing large prosthetic dyes, we achieved a high sensitivity at short spatial distances. To obtain a structural interpretation and a causal description, we compared the measurements to corresponding observables derived from both simulations and analytical models. Our integrated approach allowed us to identify key intra-protein interactions that extend beyond the MoRE region, and play an important role in NTAIL dynamics and likely also in function.

Results

Photo-induced electron transfer reveals the heteropolymeric nature of NTAIL dynamics

To characterize the dynamics of NTAIL in solution, we probed intra-molecular contact formation, using photo-induced electron transfer (PET) between tryptophan (W) and cysteine (C) residues at different positions in the sequence.3537 After exciting W to the triplet state using a nanosecond UV laser pulse, we monitored the triplet state population in time, via transient absorption measurements. As NTAIL undergoes spontaneous conformational changes, W can come into contact with C and PET can occur. During PET, the excited-state (triplet) electron is transferred from W to C, forming a W radical cation which eventually relaxes to the ground state. The relaxation time due to PET, τCW, therefore provides information on how often and quickly C and W come into contact after W excitation. If the C-W contact formation rate is of the same order of magnitude, or faster than the natural decay rate of W, τCW can be obtained from the short time behavior of the triplet state relaxation curves (Supporting Fig. S1).

The wild-type NTAIL sequence (Fig. 1A) does not contain native W or C residues. To probe the NTAIL conformational dynamics, we generated five NTAIL variants (PET variants), each bearing a unique C-W pair, to probe the five distances indicated in Fig. 1C. To minimize perturbation of the chemical properties, we substituted native tyrosine (Y) residues with W, and native serines (S) with C residues. To determine the natural decay rate of excited W at the two positions, we also generated two variants with the respective W substitution, but lacking any C substitution. The Supporting Information contains the sequences of all seven NTAIL variants (Table S1) and details about their production and purification (Supporting Methods S1).

Relaxation times τCW were estimated from multiple, independent sample preparations for each NTAIL variant, and a series of repeated measurements to account for W and C photodamage during the measurements (Supporting Methods S2). All measured relaxation curves and fitted functions to obtain τCW are shown in Fig S3S16, with the fit results summarized in Table S2.

We first performed measurements with the five PET variants of NTAIL under near physiological conditions (i.e. 15mM Tris, 150mM NaCl, 1mM TCEP, pH 7.6, in the following referred to as “reference conditions”). Figure 1D shows the τCW (black circles) estimated from these measurements as a function of the W-C sequence separation (|ij|).

To test whether the estimated τCW can be explained based on C-W sequence separation alone, we compared them to relaxation times expected from a simple homopolymer model (Fig. 1D, dashed line), where the only non-local interactions between the residues are excluded volume interactions. In this model, both the physical distance in space between C and W at positions i and j as well as the corresponding relaxation time τij scale with the sequence separation |ij|.39 The scaling relation for relaxation times τij=τr|ij|vτ depends on a prefactor τr and an exponent vτ. The exponent vτ·=1.69 (corresponding to a Flory scaling exponent v of 0.52) was estimated from small-angle X-ray scattering (SAXS) measurements of NTAIL6 based on a homopolymer model (SAW-v model40, Supporting Methods S4). The prefactor τr=0.011μs was obtained by fitting to the estimated τCW (Fig. 1D) from PET experiments.

The general trend of the estimated τCW at physiological pH (Fig. 1D, black circles) follows the simple homopolymer scaling (dashed curve), however some contacts form faster and others slower than expected. This dynamic heterogeneity (deviation from homopolymer model) observed along the sequence is most striking when comparing the C3W2 (|ij|=36) and C4W2 (|ij|=30) variants for which the C-W pairs encompass the MoRE region. These variants share the W position and differ by just six residues in sequence separation between W and C (Figs. 1A1C). According to the homopolymer model, such a difference in sequence separation would result in a τCW increase by a factor of 1.3. In contrast, the τCW estimated from relaxation curves increases by a factor of 2.8. This considerable deviation from the homopolymer behavior is primarily due to the measured 7.6 μs relaxation time for C3W2 that is substantially longer than the 4.7 μs homopolymer estimate. The fact that this anomalously slow relaxation time is not reproduced by the homopolymer model indicates that the NTAIL dynamics near the MoRE region is considerably affected by more complex heteropolymeric interactions, likely secondary structure formation, electrostatic interactions between charged residues, or solvent-induced interactions.

Charged amino acids partially account for NTAIL dynamic heterogeneity

To test whether interactions between charged residues or secondary structure formation causes the dynamic heterogeneity, we performed additional experiments at low pH and increased salt concentrations, i.e. under conditions that should modulate electrostatic interactions.

The NTAIL sequence contains 40 charged residues (13 Asp, 10 Glu, 14 Arg and 3 Lys), often clustered together (Fig. 1B). If electrostatic interactions between charged residues were markedly contributing to the observed dynamic heterogeneity, we would expect that screening the interactions by increasing the salt concentration would decrease or eliminate the heterogeneity. Indeed, increasing the salt concentration from 150mM (Fig. 1D, black circles) to 500mM (green triangles) results in a τCW decrease from 7.6 μs to 5.8 μs for C3W2, and an increase from 3.2 μs to 3.9 μs for C3W1, bringing both data points close to the homopolymer curve. This suggests that electrostatic interactions are a major contributor to the observed dynamic heterogeneity of NTAIL.

Next, we investigated why electrostatic interactions accelerate relaxation for the C3W1 variant, but slow down relaxation for the C3W2 variant. We hypothesized that changes in contact dynamics are caused by the non-uniform distribution of charged amino acids (charge patterning41,42) along the NTAIL sequence. However, the increased electrostatic screening at higher salt concentrations could also affect the stability of the transient helix in the MoRE region, and thereby modulate the C3W2 relaxation time. To address these two possibilities, we performed coarse-grained molecular dynamics (CG) simulations that allowed us to explicitly include both charge patterning and the secondary structure contributions.

For these simulations we used a modified version of the recently developed residue-based HPS model43 for IDPs. The CG model was adjusted to reproduce the experimental radius of gyration of NTAIL obtained from SAXS6 and the helical propensity of the MoRE region from NMR7. The effects of varying salt concentrations in the CG model were described by Debye-Hückel electrostatic screening44. Additional details on the CG simulations are provided in Supporting Methods S5.

To determine the effect of increased salt concentration (150 mM to 500 mM), we computed the mean distance for each pair of amino acids i and j from the CG simulations at both salt concentrations. Fig. 2A shows the difference of these means, highlighting the compaction/expansion of each NTAIL segment. In the simulations, the N-terminal part of NTAIL up to C3 expands at higher salt concentrations (red areas), whereas the C-terminal part becomes more compact (blue areas).

Figure 2.

Figure 2.

Coarse-grained (CG) simulation of NTAIL. Differences between the mean pairwise distances (Ri,j) of the CG simulations at the reference condition (0.15 M salt and pH 7.6) and simulations A) at high salt concentration (0.5 M salt, pH 7.6) or C) low pH (0.15 M salt pH 4.0), respectively. The diagonals show the charged amino acid positions with positively charged amino acids in blue and negatively charged amino acids in magenta. The MoRE is highlighted in grey. The full amino acid sequence is provided in Fig 1A. B) Ratios of C-W relaxation times due to PET, calculated from CG simulations (bars) at different conditions (high salt: green, or low pH: red), relative to the reference condition. For comparison, experimental ratios are shown where available (black squares). D) C-W relaxation times for all five NTAIL PET variants estimated from CG simulations at different conditions (reference: black, high salt: green, low pH: red). The blue triangles denote the relaxation times for the C3W2 and C4W2 variants after varying the MoRE helix content (helix− 35%, helix+: 60%). All CG relaxation times were scaled by a factor, chosen so that the same homopolymer fitting function (dashed line) of Fig 1D fit both experimental and CG data.

The positively and negatively charged amino acids are well-blended in the N-terminal region, with a close to zero net charge. The N-terminus therefore behaves like a polyampholyte, where interactions between oppositely charged residues result in compact conformations and screening of these electrostatic interactions at higher salt concentrations results in more expanded conformations. This expansion in turn increases the average distance between C3 and W1, leading to a slower expected relaxation time at high salt concentration. In contrast, the C-terminal region has more negatively charged amino acids than positively charged amino acids under physiological conditions. This imbalance leads to net repulsive electrostatic interactions that keep this region expanded. Increasing the electrostatic screening at high salt concentrations reduces the repulsion, which leads to the compaction of the C-terminus of NTAIL, a decrease of the mean C3-W2 distance, and an expected decrease in the C3W2 relaxation time.

To determine if the salt-induced changes in relaxation times can be fully described by changes in inter-residue distances due to charge patterning alone, we estimated the expected ratio of relaxation times τ/τ0 from the CG simulations at 500 mM and 150 mM NaCl concentrations, respectively. For these calculations, we assumed reaction-limited quenching kinetics, in which case, the relaxation time is inversely proportional to the probability of W and C being at a reactive distance. We used a cutoff distance of 1 nm to determine the probability of reactive conformations (further details in Supporting Methods S5). Figure 2B shows the τ/τ0 ratios upon increasing the salt concentration for each variant (green bars). A ratio larger than one indicates an increase of the relaxation time, exclusively due to the expansion of that segment. Comparison with experimental ratios (Fig. 2B, black squares) shows that the CG simulations accurately predict both the increase in C3W1 relaxation times and the decrease in C3W2 relaxation times, which indicates that charge patterning suffices to explain the salt-induced changes in NTAIL relaxation times. It also corroborates the hypothesis that electrostatic interactions between charged residues are a key factor in the observed dynamic heterogeneity in NTAIL.

To further test the effect of electrostatic interactions on the dynamic heterogeneity of NTAIL, we measured C-W relaxation times, upon decreasing the pH from 7.6 to 4.0, while maintaining a near-physiological salt concentration (150 mM) (Fig. 1D, red squares). In contrast to increasing the salt concentration which screens all electrostatic interactions, decreasing the pH only alters the charge on specific residues that titrate in the 7.6–4.0 pH range, specifically negatively charged Asp and Glu may become neutral, and neutral His become positively charged. If electrostatic interactions involving these charged residues were the main contributor to the dynamic heterogeneity observed at physiological pH, we would expect a reduction in heterogeneity when decreasing the pH to 4.0. In particular, we expect the N-terminal part, which acts as a balanced polyampholyte at pH 7.6, to become positively charged and therefore to expand at pH 4.0, even beyond what was observed at high salt concentrations. This expansion should then increase the PET relaxation times for the C1W1 and C2W1 variants. Conversely, we expect the negatively charged C-terminal part to become more neutral, resulting in more compact conformations that reduce the C3W2 and C4W2 relaxation times.

To capture the effects of lower pH in CG simulations, the charges of the residues that titrate between 7.6 and 4.0 were adjusted to match the average level of protonation at the simulated pH. To that aim, the charges assigned at pH 7.6 (−1.0 for Asp and Glu and +0 for His residues) were changed to partial charges at pH 4.0 (−0.69 for Asp, −0.36 for Glu and +0.99 for His).

Further details on charge assignments are provided in Supporting Methods S5. If electrostatic interactions involving specific charged residues drive the dynamic heterogeneity at physiological pH, lowering the pH to 4.0 should reduce this heterogeneity. The N-terminal region, a balanced polyampholyte at pH 7.6, is expected to become positively charged and expand, while the negatively charged C-terminal region should become more neutral and compact. The CG simulations with simple fixed-charge adjustment at low pH are consistent with the expected behaviors described above, i.e. the N terminal region expands and the C terminal region becomes more compact (Fig. 2C), which is also reflected in the calculated relative relaxation times (red bars Fig. 2B).

The C-W relaxation times of all five PET variants estimated from measurements at pH 4.0 are closer to the homopolymer curve (Fig. 1D, red squares) than those at pH 7.6. Lowering the pH also reduces the observed dynamic heterogeneity between the C3W2 and C4W2 variants, with the ratio of relaxation times reduced from 2.8 at pH 7.6 to 1.5 at pH 4.0, which is closer to the homopolymer prediction of 1.3. However, in contrast to the reduction of the C-W relaxation time expected from compaction based on charge patterning and predicted by CG simulation, the measured relaxation time for C4W2 increases slightly upon lowering the pH (Fig. 1D).

In Fig. 2B the ratios of relaxation times at pH 4.0 and pH 7.6 estimated from the CG simulations (red bars) were compared to those estimated from PET experiments (black squares). The CG model correctly captures the trend of C1W1 and C3W2, but not that of C4W2 and C3W1. Further, the measured increase of τCW for C4W2 is not only contrary to the CG simulation results, but also contrary to the expectation based on charge patterning, suggesting that other unaccounted factors influence the relaxation times. In the following, we will consider three factors that can affect relaxation times in a pH dependent manner: the MoRE helicity, W to C electron transfer efficiency, and residue specific changes in protonation equilibria.

pH-dependence of MoRE helix content.

The CG model suggests limited difference in helical propensity upon changing pH (Supporting Fig. S20). We further measured and compared the far-ultraviolet circular dichroism (CD) spectra of the NTAIL variant Y518W (W2) at pH 7.6 (Fig 1C, inset, black curve) and 4.0 (red curve). CD spectroscopy in this wavelength range of 180–260 nm is highly sensitive to the presence of α-helices and the two spectra appear to be very close to each other, indicating that the helical content of NTAIL is not affected significantly by pH. In addition, analysis of the measured spectra by the SESCA software45 (Supporting Methods S13) predicts at most 8% global helix content, which is compatible with the NMR estimates used to set the MoRE helicity in CG simulations. Together, these results suggest the MoRE helix content is not affected by the change in pH.

Changes in electron transfer efficiency.

Because PET involves formation of W radical cations, the pH may directly affect the electron transfer process in solution, independent of pH-induced changes in the protein conformational ensemble and dynamics. To address this possibility, we performed a bimolecular quenching study between N-acetyl-tryptophan-amide (NATA) and cysteine molecules, freely diffusing in solution, at pH 7.6 and 4.0 (Supporting Methods S2). In these experiments the bimolecular quenching rate was determined by measuring the W triplet relaxation times in samples at fixed NATA concentration and increasing cysteine concentrations. These measurements showed that the C-W bimolecular quenching rate (determined from the slopes of Stern-Vollmer plots) at pH 7.6 and 4.0 are very similar (Fig. S19). This result shows that the efficiency of electron transfer from the excited state W to C, under the conditions of our PET experiments is not directly affected by pH, and the pH dependence of τCW in NTAIL is due to changes in the protein structure and dynamics.

Residue-specific protonation.

NTAIL is rich in Asp and Glu, amino acids that reach protonation equilibrium close to pH 4.0. Under these conditions, the protonation state of individual acidic residues can be influenced, e.g., by the spatial vicinity of similarly or oppositely charged residues, resulting in a dynamic, conformation-dependent charge distribution.46 This charge distribution would also allow electrostatic interactions to stabilize certain conformational states, increase or decrease the observed for individual NTAIL variants compared to those predicted by our fixed-charge CG simulations.

Dynamic heterogeneity at the NTAIL C-terminus

Although dynamic protonation may explain deviations from CG and homopolymer models, this effect is an unlikely explanation for the dynamic heterogeneity between C3W2 and C4W2 under reference conditions at pH 7.6. To address this heterogeneity with CG simulations, we assumed these processes are reaction limited, and that τp the proportionality between C-W contact probabilities (Prij<r0) and corresponding τCW relaxation times are the same for all PET variants, and thus τij=τp/Prij<r0. Under these assumptions, it is possible to compare relaxation times of different C-W pairs obtained from CG simulations relative to one another and address the dynamic heterogeneity at the NTAIL C-terminus.

To this aim, we fitted a homopolymer model to these CG-derived relaxation times at the reference condition, and chose τp such that the scaling equation is identical to the one obtained from experimental measurements in Fig. 1D. The resulting relaxation times from CG simulations for all three conditions (reference, high salt, and low pH) are shown in Fig. 2D. The figure shows that all CG-derived relaxation times at the reference condition (black circles) follow homopolymer scaling (dashed black line), with only small deviations for individual CW pairs. Although the CG simulations accurately predict many of the changes upon increasing salt concentrations or lowering the pH, they do not show either the considerable slowdown for the C3W2 relaxation time, or the pronounced dynamic heterogeneity measured at pH 7.6 between C3W2 and C4W2 (ratio 2.8). In fact, the CG-derived τC3W2/τC4W2 ratio of 1.6 is close to the 1.3 ratio predicted by homopolymer models, suggesting that the CG simulations do not capture the main cause of the dynamic heterogeneity observed in PET measurements.

An additional reason why the CG simulations may not have captured the C3W2/C4W2 dynamic heterogeneity is that we underestimated or overestimated the MoRE helix propensity (~45%) under the conditions of PET experiments. This can alter the CG-derived heterogeneity if the MoRE helix in its folded state affects C3W2 and C4W2 relaxation differently. To test this hypothesis, we performed additional CG simulations in which we altered dihedral potentials that control the MoRE helicity and computed the τCW for C3W2 and C4W2 variants at high (60%) and low (35%) helix contents (blue triangles on Fig. 2D). Relaxation times from these simulations show that considerable changes in the helix content can indeed shift the relaxation times obtained from CG simulations, but this shift affects both C3W2 and C4W2 similarly. Therefore, even a significant error in the imposed MoRE helicity would have only a small (±0.1) effect on τC3W2/τC4W2 ratio, certainly not sufficient to explain the dynamic heterogeneity observed in PET experiments.

Overall, our results show that NTAIL dynamics is heteropolymeric in nature, especially in the region encompassing the MoRE, with electrostatic interactions playing an important role. It appears that the dynamic heterogeneity of NTAIL cannot be explained by a simple polymer scaling model (Fig. 1D) or by CG models including helix formation within the MoRE and explicit interactions between charged amino acids (Fig. 2). These observations suggest that the cause of the dynamic heterogeneity is either side-chain specific transient interactions, such as salt-bridges and hydrogen bonds47, or interactions between the protein and solvent. To study the effects of these interactions on NTAIL dynamics in more detail, we turned to all-atom explicit-solvent molecular dynamics simulations.

Disordered NTAIL is characterized by a conformational ensemble with distinct states

In addition to specific interactions between amino acids and protein-solvent interactions, all-atom simulations include the backbone and side-chain dynamics required to accurately compute the contact formation rates. The relaxation times τCW estimated from these contact formation rates can be compared to τCW derived from PET experiments. However, this increased level of detail comes with two drawbacks; considerably higher computational costs that limit conformational sampling and the need for a larger number of accurate interaction parameters (force field). IDP simulations are particularly sensitive to variations in the used force field4851, and force field accuracy for IDPs often varies in a system-dependent manner52.

To maximize conformational sampling, we limited all-atom simulations to the reference conditions (pH 7.6 at 150 mM NaCl concentration) and focused on the C3W2 and C4W2 variants, which showed large dynamic heterogeneity (Fig. 1D). Further, to ensure the reliability of all-atom simulations, we first validated eight force fields, characterized by different combinations of protein and water parameters, against independent experimental SAXS, CD, and NMR6,16 data on the wild-type (WT) NTAIL sequence (shown in Supporting Table S3). To evaluate these force fields, we first performed simulations of WT NTAIL using each force field, and then compared the resulting trajectories to the SAXS data, which reports on protein compactness, using the software CRYSOL53. Next, we compared the trajectories to CD and NMR chemical shift measurements, which report on secondary structure propensities, using SESCA45, and SPARTA+54. For further details of the force field validation, see Supporting Methods S8.

The two force fields that reproduced the WT validation data most accurately were the CHARMM36m (C36M)51 force field with optimal point charge (OPC) water model55, and Amber99SB-disp (A99SB-d) force field49. Both of these force fields produced extended NTAIL ensembles with transiently forming helices in the MoRE region, that were in good agreement with the available experimental data (Supporting Table S3), and suggested helix propensities similar to those reported in computational studies of MoRE region19,26 (Supporting Methods S23).

The two best force fields were used to perform longer all-atom simulations of the C3W2 and C4W2 variants. As a second force field validation step (Supporting Table S4), the variant trajectories were tested against PET relaxation measurements of the same variants, and the CD spectrum of the W2 variant (Fig. S21B). Simulations of both force fields estimated C3W2 and C4W2 relaxation times showing considerable dynamic heterogeneity. However, the C36M-OPC force field was the only parameter set that reproduced measurements of both WT NTAIL and the two variants within the experimental and computational uncertainty, including the two measured W relaxation curves (Supporting Fig. S22). Thus C36M-OPC trajectories were used for all subsequent analysis. The relaxation times τCW estimated from these simulations were 2.0 ± 0.5 μs for C4W2 and 5.3 ± 3.3 μs for C3W2, in good agreement with the corresponding measured values of 2.3 μs and 7.6 μs. Although the uncertainty on the C3W2 relaxation time is large due to limited sampling, the ratio of the relaxation times estimated from the simulations is 2.7 ± 1.8 close to the experimentally determined ratio of 2.8. Together, these results suggest that the all-atom simulations accurately describe the NTAIL dynamics causing the C3W2/C4W2 dynamic heterogeneity.

To study the origin of the dynamic heterogeneity in detail, we computed the free energy landscape of the NTAIL variants along the C-W distances for positions C3W2 and C4W2. Variants C3W2 and C4W2 only differ in the exchange of an oxygen and a sulfur atom between two nearby residues at position 482 (C3) and 488 (C4), respectively. We therefore assumed that the free-energy landscape of the two variants is similar and combined their trajectories for computing it. The C-W distances were defined as the distance between the gamma sulphur of C (or gamma oxygen of S) and the nearest heavy atom of the W side-chain ring. These serve as intuitive reaction coordinates for describing C-W contact dynamics in the NTAIL C-terminal region.

Unexpectedly, the computed free-energy landscape (Fig. 3) is characterized by four local minima corresponding to four NTAIL conformational states (S1 to S4). Approximately 60% of the sampled NTAIL conformations are within the respective energy wells, including almost all conformations with C-W contacts necessary for PET. S1 corresponds to the conformational state around the absolute free-energy minimum and is occupied for approximately 24% of the simulation time. S2 and S3 are less populated conformational states, occupied for 14% and 8% of the simulation time, respectively. S4 is the least populated state, with approximately 2% of conformations. The landscape shows only small barriers (< 1 kT) separating S1, S2, and S3, and a low free-energy barrier between S1 and S4 (< 2.0 kT). This shallow and accessible free energy landscape suggests nearly-free diffusion between states, which is in line with previous SAXS results suggesting NTAIL to be a pre-molten globule with residual structure6.

Figure 3.

Figure 3.

Conformational states revealed by all-atom MD simulations. The free-energy landscape of sampled NTAIL conformations along the calculated C-W distances from C36M-OPC MD simulations. Contour lines denote a 0.3 kT difference in the estimated free energy landscape. Conformational states are highlighted by colored labels and rectangles. The population of conformations within each rectangle is shown in the legend. Red dashed lines indicate a contact distance of 0.4 nm. Conformations below these cutoffs would allow quenching of the W triplet state, due to electron transfer to C, for the corresponding variant (C4W2 or C3W2).

The four conformational states influence C-W relaxation times as well. PET cannot occur in the two most populated states (S1 and S2), because the distances between electron acceptors C4 or C3 and electron donor W2 are more than 2 nm. In state S3, C4-W2 distances below 0.4 nm are observed, allowing PET in the C4W2 variant, while in state S4, C3-W2 distances below 0.4 nm are reached, which allows PET in the C3W2 variant. Further, from the most populated S1 state the free-energy barrier that NTAIL has to cross to reach S4 is about 0.9 kT larger than the barrier to reach S3. This suggests that S3 is more easily accessible than S4, with an approximately 2.5-fold difference between the respective transition rates. This result is consistent with and provides a possible explanation for the measured 2.8 ratio between the C3W2 and C4W2 relaxation times.

Non-local NTAIL interactions explain dynamic heterogeneity

To determine the molecular basis of the observed dynamic heterogeneity, we investigated which specific interactions stabilize the four conformational states identified above. To this aim, Fig. 4 shows intramolecular contact probability maps for each conformational state, with backbone (Cα) contacts shown in yellow to green. Because both the pH and salt concentration dependence of PET experiments and CG simulations indicated an important role of electrostatic interactions in NTAIL dynamics, we also calculated contact probability maps for salt bridges between charged amino acids shown in purple to blue. For the sake of clarity when discussing the contacts, we defined five regions along the NTAIL sequence (Fig. 4 bottom, color-coded and labelled A-E). Regions A-C are subdivisions of the polyampholyte N-terminal part identified in our CG simulations, regions D and E constitute the negatively charged C-terminal part, with the MoRE located in region D.

Figure 4.

Figure 4.

Contact maps and representative conformations in NTAIL conformational states. For each conformational state (S1-S4, shown in Fig.3), a representative conformation (left) and contact probability map (right) is shown. Each map shows Cα contact probabilities (above the diagonal, yellow to green) and salt bridge probabilities (above and below the diagonal, purple to blue). Cysteine acceptor positions C3 and C4 (residues 482 and 488) as well as tryptophan donor position W2 (residue 518) are indicated by cyan and orange dashed lines, respectively. Potential C-W contacts are indicated by black circles. The boundaries of identified interaction regions A-E and their color-coding for representative conformations are shown in the bottom-left side of the figure. The same regions are separated by black dotted lines on the contacts maps. The most prominent interactions in each state are encircled. The pixel size of salt-bridge interactions was increased for visibility, using a Bessel interpolation scheme. On the left side, the backbone trace of a representative conformation from each state is shown in cartoon representation, colored according to regions. The donor and acceptor atoms are highlighted as orange and cyan spheres, respectively. Residues with prominent interactions in the conformational states are labelled and shown in stick representation. Individual contact maps in high resolution, as well as the average contact map for all NTAIL simulations are shown in Figs. S23S27.

The contact maps (Fig. 4, right) clearly show that the four states are stabilized by distinct sets of local (within a region) and non-local (between regions) interactions, although some interactions occur in several conformational states. The maps also provide information on the secondary and tertiary structure preferences of NTAIL in these conformational states. Helical secondary structure elements and tight turns appear on contact maps as continuous sets of interactions, parallel and close (within 5 residues) to the diagonal. β-strands, β-hairpins or longer loops appear as sets of interactions that are further away and either parallel or perpendicular to the diagonal. Disordered regions typically have few or sporadic interactions.

The four conformational states show markedly different structures (Fig. 4, left), including the MoRE in region D. In state S1, region D is largely non-helical but forms several transient local salt bridges. Several non-local interactions (encircled in green) may prevent donor-acceptor contacts for both variants (C3W2 and C4W2). S1 conformations are also stabilized by salt bridges formed between regions B and E. Salt bridges involving K441, R444 (region B) and E507 and D511 (region E) are particularly frequent and probably play an important role in stabilizing this conformational state.

State S2 includes conformations where region D is non-helical and conformations with a folded α-helix (residues 492–502). The helical structures appear on the contact map as an interaction patch close to the diagonal of region D (encircled in orange). Non-local interactions in S2 appear to be less stable than in S1 and include contacts between regions A and C, C and D, as well as B and D. Stable interactions in S2 also involve two salt bridges between R413 and D476 (regions A and C) and K441 and E449 (local in region B). Many sporadic interactions (both backbone contacts and salt bridges) were observed within the N-terminus (between regions A and B). Taken together, the interactions suggest that S2 consists of multiple conformational substates and most interactions are linking central regions (B and C) with both termini (A, D, and E).

Compared to S1 and S2, the conformations of S3 comprise fewer non-local interactions and are instead characterized by two very stable local interaction sites. These local interactions correspond to a shorter helix between residues 492–497 present in all S3 conformations and to a stable hairpin or zipper in region B, stabilized by a salt bridge (K441-E449) observed previously in S2. Further, there are two sets of notable non-local interactions. First, a patch of interactions between regions D and E folds the C-terminus into a loop, stabilized by the salt bridge between R490 and D520. These interactions also bring the electron donor W2 (518) very close to the potential acceptor at position C4 (488), thus promoting electron transfer in the C4W2 variant. At the same time, the zipper in region B brings the N-terminal close to regions C and D, where a salt bridge between K405 and D484 may prevent electron transfer to position C3 (482) in the C3W2 variant.

The S4 state allows electron transfer to position C3 (482) in the C3W2 variant. It is structurally very different from states S1-S3. The most defining structural feature of this state is a long loop reminiscent of an anti-parallel β-sheet stretching over the entire C-terminus (regions D and E). This loop is stabilized by interactions between regions C and E including a salt bridge between R463 and D513. The N-terminus of NTAIL in this conformational state has only weaker interactions, mostly limited to contacts between regions A and C as well as between regions B and C.

To determine if frequent contacts observed in one of the four conformational states (Fig. 4) are specific to that state and if they are correlated with other contacts, including C-W contact formation for the C3W2 and C4W2 variants, we applied normalized pointwise mutual information analysis (NPMI, see Supporting Methods S22).

The NPMI analysis revealed that many salt bridges between regions B, D, and E occur selectively in their respective conformational states. These stable salt bridges often co-occur with clusters of backbone contacts (encircled in Fig. 4), possibly nucleating the interaction clusters that stabilize the states.

Further, all frequent contacts observed in the S1 state were anti-correlated with C-W contact formation (Table S5), indicating that the specific interactions that stabilize the S1 state would increase the expected relaxation times compared to the flat free-energy landscape of a homopolymer model. In contrast, all frequent contacts of states S3 and S4 are positively correlated with C-W contact formation for variants C4W2 and C3W2, respectively. Thus, interactions in these two states selectively reduce the relaxation times of either C4W2 or C3W2 variants, contributing to the dynamic heterogeneity observed in PET experiments.

Finally, the same charged residues (e.g., K441, R497, and D501) in regions B, D, and E form salt-bridges with different residues in different conformational states (Table S5). Switching between these sets of interactions is necessary to transition between the most populated conformational states, and therefore likely define the free-energy barriers between the states (Fig. 3). Further, a conformational switch based on these competing intramolecular NTAIL interactions of region B may have a functional role, by regulating the helical propensity of the MoRE in region D, which in turn modulates affinity for the measles phosphoprotein.

Functional relevance of non-local interactions is supported by coevolutionary analysis

Having identified the most important interactions that dictate NTAIL structure and dynamics (Fig. 4), we asked whether they also have a functional role. To this aim, we performed a coevolutionary analysis of NTAIL using EVcouplings56, which identifies and quantifies correlated mutations of residue pairs within the sequence of homologous proteins from different species or strains, and which often indicates pairwise-interactions of functional importance. A total of 12,072 sequences similar to NTAIL have been aligned and selected for this analysis (Supporting Methods S21).

Most non-local residue pairs with positive EVcouplings scores connect regions B and D (Fig. S28), suggesting evolutionarily conserved interactions between these two regions. Here we specifically highlight the four non-local residue pairs with the highest EVcoupling scores (Fig. S28, top left, Table S5). These pairs include hydrophobic contacts involving three leucine residues (L491, L496 and L498) interacting with S443 and A447, and one salt bridge between R444 and D493. Correlated mutations of residues in the distant regions B and D suggest a biologically relevant, functional interaction between these two regions of NTAIL. Given that both L491 and L498 are located in the MoRE region (in region D) and are directly involved in binding to PXD, it is possible that their interactions with residues of region B serve the function of regulating binding. We note that the contact analysis of our simulations does not identify these specific coevolving residue pairs as frequent interactions. However, all-atom simulations independently identify non-local interactions between regions B and D (Fig. 4, S2 state), including A447 contacts with PXD binding residues L498 and M501. These differences between all-atom simulations and the co-evolution analysis may be due to insufficient sampling or the lack of biological context in all-atom simulations (i.e. the absence of the NCORE domain or PXD itself).

Discussion

To understand the role of disordered regions flanking the MoRE on either side in NTAIL dynamics and conformational preferences, we probed non-local interactions of the full-length protein in solution, combining photo-induced electron transfer experiments, analytical polymer models, molecular dynamics simulations, as well as coevoluionary analysis. With PET experiments, we measured and compared contact formation times between tryptophan and cysteine residue-pairs introduced in different regions of the NTAIL sequence, under varying solution conditions. The experiments revealed a pronounced dynamic heterogeneity within the protein (Fig. 1) under physiological conditions, especially at the NTAIL C-terminal region, close to the MoRE, which is involved in binding to PXD. PET experiments also showed a reduced heterogeneity under conditions that weaken electrostatic interactions, such as increased salt concentration or lower pH, pointing to an important role of electrostatic interactions between charged residues in the dynamics of NTAIL. Based on our molecular dynamics simulations, we attribute this heterogeneity to salt-bridge induced non-local interactions between specific regions of NTAIL.

Interpretation of the general role of electrostatics

This finding is supported and explained in structural terms by our simulations (both coarse-grained and all-atom). Even without including atomic-level details, our CG simulations suggest that when increasing the electrostatic screening or changing the charge of acidic residues to emulate pH related protonation, the N-terminal region expands while the C-terminal region compacts. In addition, CG simulations performed with different helical propensities in the MoRE indicate that changes in helicity of the MoRE do not give rise to the dynamic heterogeneity within the NTAIL C-terminal region (i.e. C3W2 and C4W2). These results indicate that side-chain specific transient interactions, not described by the CG model, contribute to the large difference between the C3W2 and C4W2 relaxation times.

Specific non-local interactions dictate NTAIL conformation and dynamics

Our all-atom simulations, performed with a carefully validated force-field, describe interactions with atomic-level detail. These simulations reproduce the dynamic heterogeneity of C-W pairs encompassing the NTAIL MoRE (C3W2 compared to C4W2). We found that, while NTAIL is clearly disordered, a few key non-local interactions stabilize four distinct conformational states which in total represent 48% of the structural ensemble. The two most populated states do not allow PET for either C3W2 or C4W2 variants, while the third and fourth states allow PET for the C4W2 and C3W2 variants, respectively. The free energy landscape of NTAIL (Fig. 3) reveals rapid interconversions between the three most populated states. In contrast, structural rearrangements required to reach the C3W2 contact state are slower due to a small, though comparatively higher free-energy barrier (~4.4 kJ/mol) under physiological conditions. This free-energy barrier and the lower equilibrium population of the C3W2 contact state contributes to the dynamic heterogeneity of the NTAIL C-terminal region.

Analysis of the state-specific contacts (Fig. 4) reveals that competing clusters of interactions stabilize the four conformational states. These interaction clusters involve (and are possibly nucleated by) stable salt bridges between charged amino acid side chains. Further, focusing on the two most populated conformational states S1 and S2, our all-atom simulations show that specific charged residues in region B (residues 435–451) can form either non-local salt bridges with C-terminal regions D-E (484–525, close to the MoRE), or local salt bridges with other region B residues. Switching between these sets of interactions (as sketched in Fig. 5) is necessary for NTAIL to transition between the two dominant states, affecting its large-scale dynamics.

Figure 5.

Figure 5.

Schematic representation of the two major NTAIL conformational states (S1 and S2). Different regions of NTAIL are color-coded according to Figure 4. Non-local interactions between the regions are represented as dashed lines. Key interactions between regions are highlighted by orange shading. The helical binding site of the phosphoprotein X-domain (PXD) in region D is marked by an orange arrow.

Co-evolutionary analysis independently shows that correlated mutations across regions were only found between regions B and D. Given the distance in sequence between these two regions, this not only corroborates the presence of non-local interactions between region B and D, but also suggests that they play a functionally important role.

In summary, our simulations and coevolutionary analysis independently highlighted region B of NTAIL as a potential allosteric interaction partner to region D, which includes the binding site (MoRE) for the phosphoprotein X-domain (PXD). Additionally, fluorescence-based binding kinetics experiments on truncated NTAIL variants23, indicated that intramolecular interactions of NTAIL involving residues 435–451 (region B) have a significant impact on PXD binding in vitro, despite being located far away in sequence from the MoRE.

Identified interactions in the context of NTAIL to PXD binding

We hypothesize that the intramolecular NTAIL interactions observed in our simulations not only govern the NTAIL dynamics but also play a role in regulating PXD binding to the MoRE. Mutational studies by Bignon et al.57 have shown that the helix content of the MoRE correlates with the binding affinity, and that N-terminal truncation of NTAIL increases its affinity for PXD22,24,27. Similarly, Gruet et al.27 have identified mutations in several regions of the NTAIL sequence, on either side of the MoRE, which increase PXD binding affinity. The crystallographic structure of the NTAIL MoRE bound to PXD9 (PDB: 1t6o, Fig. 6B) shows that the binding interface consists mainly of a hydrophobic edge along the MoRE helix (region D), between residues S491 and M501 (hereafter D1-site, shown in dark green), which interacts via intermolecular contacts with the α2-α3 face of the PXD triple helix (Fig. 6A). The other side of the MoRE helix is a hydrophilic edge formed by charged and polar residues (D2-site, dark red) between Q486 and R497. The NTAIL-PXD complex is further stabilized by intermolecular interactions involving R497 and salt bridges formed by D487 and R490 which are located at the beginning of the MoRE helix between the two edges (D3-site, light green)27. Many of these key amino acids are involved in intramolecular NTAIL interactions in one or more conformational states, as shown by the contact probabilities obtained from all-atom simulations (Fig. 4).

Figure. 6.

Figure. 6.

Speculated biological context of non-local NTAIL interactions. (A) NTAIL MoRE residues in region D constituting the hydrophobic PXD-binding interface are highlighted (D1 site, dark green) along with charged residues possibly involved in intramolecular NTAIL interactions (D2 site, dark red), and residues possibly involved in both intra- and intermolecular interaction (D3 site, light green). (B) Cartoon representation of the PXD-MoRE complex based on the available crystal structure (PDB: 1t6o). (C) During the initial steps of the virus replication the MeV nucleocapsid protein (N, blue) recruits the large polymerase (L, green), by binding to the phosphoprotein (P, orange). Here NTAIL is extended. (D) When Brownian motion allows L to approach the RNA (red) in the nucleocapsid groove, NTAIL adopts a hairpin conformation, which promotes intramolecular interactions between the MoRE and the N-terminal region of NTAIL. (E) Once the contact between nucleocapsid and L is established, additional complexes can be formed between free PXD and downstream N monomers. Further, intramolecular NTAIL interactions help release bound P to allow polymerase progression, by weakening the binding to PXD when L is nearby. Numbers (1–4) on PXD denote which step individual domains are in a “cartwheeling” L progression mechanism. In panels C, D, and E, only 1–3 NTAIL domains have been depicted for the sake of clarity. The RNA molecule is shown as a solid line on its first helix turn around the nucleocapsid to indicate that it is more solvent exposed, and as a dashed line on downstream turns to highlight that it is partially buried in the RNA-binding ridge.

Based on these interactions, we propose two independent mechanisms that provide a possible explanation for the observed changes in PXD binding affinity upon NTAIL mutation or truncation.

Direct competition for MoRE:

In all-atom simulations, the D1-site residues (Fig. 6A, dark green) form intramolecular interactions with residues of regions B and C in state S2 as well as with residues of region E in state S3 (see Supporting Table S7), both of which are states with considerable helical propensity of the MoRE region. Assuming that a folded MoRE helix is a necessary part of the biologically active NTAIL conformation, these intramolecular interactions would directly compete with the intermolecular interactions with PXD.

First, the intramolecular interactions would reduce the free helix population of the MoRE, thereby decreasing the binding rate of PXD. Second, the intramolecular interactions may also displace PXD from the binding site without exposing the hydrophobic D1 site to water, thereby lowering the PXD unbinding activation barrier and increasing the unbinding rate. Both effects would regulate MoRE-PXD binding by reducing the binding affinity, and explain why the binding affinity increases when flanking regions of the MoRE are removed27,57.

Indirect allosteric regulation:

Helical conformations of the NTAIL MoRE appear to have higher affinity for PXD57. Therefore, changing the population of helical MoRE conformations via intramolecular interactions would also affect the MoRE-PXD binding affinity. This mechanism is similar to the ensemble allosteric model for IDPs described by White et al29. Specifically, intramolecular interactions that stabilize non-helical MoRE conformations, such as the contacts seen in state S1, would decrease the population of helical conformations, reducing both NTAIL binding affinity and the rate of binding to PXD. The observed contacts (Fig. 4) suggest that the ratio of states S1 and S2 is controlled by the competing interactions of region B, which acts as a conformational switch. Specifically, in state S1, the interactions between regions B and E stabilize a loop-like, non-helical conformation of the NTAIL C-terminus, which is likely binding-incompetent towards PXD. In contrast, in state S2, region B residues engage in local interactions, allowing the NTAIL C-terminus to form the high-affinity helical structure in the MoRE at region D.

Potential functional role of NTAIL intramolecular interactions in viral replication

Non-local interactions between regions B and D/E of NTAIL may also play a functional role in RNA polymerization and viral replication. In the following we present three established mechanisms where intramolecular interactions between the B and D regions of NTAIL may contribute to optimizing the binding kinetics for viral replication.

Tether shortening:

Initial polymerase complex recruitment likely involves the formation of one or more disordered tethers between the polymerase complex and the nucleocapsid. By limiting the volume available for diffusion, tethering increases the local concentration of the polymerase complex around the nucleocapsid, thereby increasing the rate of transcription/replication initiation. There are two experimentally established tethers between N and P. The first tether consists of the disordered N-terminal tail of P (1–304, PTAIL) binding to two opposite sides of NCORE, and its primary function appears to be the prevention of nucleocapsid association in the absence of the viral RNA58. The second, tether is formed by regions B and C of NTAIL and the C-terminal domain (PLOOP, 376–459) of P, through coupled binding and folding between the MoRE of N and PXD (Fig. 6C). This tether regulates L progression on the nucleocapsid.59 The two tethers may coexist and in fact work in concert to form N-P condensates that serve as the basis of viral factories within the host cell60,61, and to securely attach the polymerase complex to the nucleocapsid during transcription/replication initiation11.

During replication initiation, it is preferable to have long tethers to recruit polymerase complexes from a larger volume around the nucleocapsid, but later, when L has to be guided to the 3’ end of the RNA, shorter tether lengths may be preferable. Intramolecular NTAIL interactions between the MoRE D2-site and region B would shorten the tether by 45 amino acids (Fig. 6D), further reducing the volume available to the polymerase complex, and increase the rate of initiation of genome transcription and/or replication.

PXD unbinding during polymerase progression:

Sourimant et al.10 carried out recombinant minigenome bioactivity assays found that polymerase progression is slowed down considerably by a deletion of regions B and C of NTAIL and proposed a mechanism for polymerase progression. This mechanism is based on a newly discovered binding site on NCORE for the α1-α2 face of PXD, at which compensatory point mutations could weaken PXD-NCORE binding and restore bioactivity. Briefly, the mechanism suggests that the L-P polymerase complex is tethered to the nucleocapsid by forming the NTAIL MoRE− PXD complex. Then, the PXD-MoRE complex binds the recently discovered NCORE site. PXD binding to this NCORE site likely facilitates PXD-MoRE unbinding, possibly after compaction of the NTAIL loop to allow polymerase passage. PXD unbinding in turn allows subsequent binding to MoRE of the downstream N monomer, and polymerase translocation.

The mechanism proposed by Sourimant et al. provides previously unprecedented molecular insight into the measles virus replication. However, it mostly considers the deleted regions of NTAIL as passive linkers or “roadblocks” for L, and it does not give an in-depth explanation on how weakening the NCORE-PXD binding affinity compensates for reduced bioactivity caused by the deleted NTAIL segments. Our current results and previous binding efficiency studies on truncated NTAIL constructs22 suggest that non-local interactions between MoRE and NTAIL regions B and C can in fact actively compete with PXD for its binding site and may do so in the context of a MoRE-PXD-NCORE ternary complex as well. This implies that deleting regions B and C does not just shorten NTAIL but also eliminates intramolecular interactions that promote MoRE-PXD unbinding. This would explain why the deletion hampers viral replication, why mutations at the NCORE interface compensate for the deletion, and provides a possible reason for region B residues to co-evolve with the MoRE (Table S5). In addition, based on the available cryo-EM structure of the nucleocapsid1 (PDB: 4uft), region B would emerge from the RNA binding ridge close to the NCORE binding site, with sufficient space for region B to fold into a loop and interact with the MoRE region (Fig. S24).

Polymerase progression by cartwheeling:

Our findings are also compatible with a modified version of the cartwheeling mechanism62 for polymerase progression (Fig. 6E). Here, P is recruited (Fig. 6E, step 1) by NTAIL and bound to the NCORE binding site (step 2) of the same monomer as described by Sourimant et al.10. However, in this mechanism PXD remains bound to the same N monomer until the L polymerase approaches. Then, due to steric hindrance from L or conformational changes at the N-N interface, NTAIL folds back onto the nucleocapsid (step 3) in a conformation that allows region B to weaken PXD-MoRE interactions. This, in turn, allows PXD to unbind and be transferred to its binding site on L (step 4) as L progresses. Subsequently, PXD is either released to the solvent or directly binds to the NTAIL of a downstream N monomer in the next cartwheeling cycle.

As discussed above, evidence suggests intramolecular NTAIL interactions weaken PXD binding. Tight binding to PXD is beneficial for tethering the viral polymerase to the nucleocapsid during polymerase recruitment and progression, and thereby ensures that the transcription and replication of the genome is completed. However, unbinding of individual PXD-NTAIL complexes is required for polymerase progression during viral replication63. Therefore, MeV and related viruses may be at an evolutionary disadvantage if the binding between PXD and NTAIL is so tight that it limits the replication rate, as supported by mutational studies of PXD64. It is thus tempting to hypothesize that a possible functional role of non-local intra-protein NTAIL interactions is to provide a conformation-dependent optimization of PXD binding affinity.

Conclusions

Our combined experimental and simulation approach shows that NTAIL is clearly disordered, contains very little secondary structure, and shows nearly diffusive dynamics. However, it still displays some structural features which cause its dynamics to deviate significantly from that of a homopolymer. In particular, we identified several key transient interactions between disordered regions distant in sequence as the main reason for this deviation. These interactions affect the overall conformation and dynamics of NTAIL in solution. Interestingly, similar interactions also emerge from our independent co-evolutionary analysis, corroborating our simulation results and suggesting that they not only govern the conformational dynamics of the essential NTAIL domain, but also are functionally important. We therefore propose possible mechanisms by which these non-local interactions regulate binding to the phosphoprotein X domain (PXD), and consequently, recruitment and progression of the polymerase complex onto the nucleocapsid template. Our results for NTAIL corroborate the importance of flanking regions in IDPs that bind via short molecular recognition elements (MoREs)64, and suggest a regulatory role of regions that may be far in the sequence from MoREs. These non-local interactions within NTAIL can regulate binding via an ensemble allosteric model29,30 without requiring the binding of a third molecule. Further studies targeting the NTAIL-PXD interaction may provide additional insight into the underlying mechanisms and effects of these interactions.

It is plausible that similar mechanisms are also at work for other IDPs that share sequence characteristics, such as charge patterning, with NTAIL. Specifically, non-local intra-protein interactions in IDPs may regulate the conformational preferences and dynamics of both the MoRE and of the entire protein, either via direct competition or indirectly by shifting the population of free-energy minima.

Methods

Protein expression, purification, and sample preparation.

Wild type (WT) NTAIL and the seven variants listed in Supporting Table S1, were expressed in E. coli, and further purified as described in detail in Supporting Methods S1. Briefly, after expression of tagged protein in E. coli, cell pellets were resuspended in 8 M urea, 50 mM Tris pH8, 0.3 M NaCl and frozen. After thawing, the solution was sonicated, spun, and tagged proteins were purified using Ni-Sepharose fast flow beads (Cytiva). After dialyzing the eluent in 50 mM Tris pH8, 0.3 M NaCl, TEV protease was used to remove tags from the desired protein. TEV protease as well as uncut tagged proteins were eventually removed using Ni-sepharose beads. Purified NTAIL proteins, free of histidine tags, were dialyzed against 15 mM Tris, 150 mM NaCl, 1 mM TCEP buffer at neutral pH, checked using SDS-PAGE, UV absorption, and far-UV circular dichroism (CD), and frozen for shipping. After shipping, samples were further purified via HPLC (semipreparative Vydac C18 column) and lyophilized. Before PET experiments, the lyophilized protein was dissolved directly into filtered buffers (for reference conditions: 15 mM Tris, 150 mM NaCl, 1 mM TCEP pH 7.6; for low pH measurements: 20 mM NaAc, 150 mM NaCl, 1 mM TCEP pH 4.0; for high salt concentrations: 15 mM Tris, 500 mM NaCl, 1mM TCEP pH 7.6). Protein concentration was adjusted to ~100 μM as determined by UV-absorbance at 280 nm (extinction coefficient 6990 cm−1 M−1). Samples (~350 μL) were placed in 5×10×30(h) mm Spectrocell quartz gas tight cuvettes with screwcap (equipped with Teflon coated silicon membrane), and bubbled with USP grade nitrous oxide for at least one hour, to reduce the concentration of dissolved oxygen (a quencher of excited tryptophan triplet state), and to introduce solvated electron scavenger (to prevent reactions with electrons, produced by water decomposition under UV laser pulses).

Photo-induced electron transfer (PET) experiments.

To probe intra-molecular contact formation, we used a technique based on photo-induced electron transfer (PET) between a tryptophan (W) and a cysteine (C)3436 placed at different positions within the sequence (sequences in Supporting Table S1). Details are described in Supporting Methods S2. Briefly, we used a homebuilt nanosecond transient absorption apparatus to excite the W to the triplet state, and to monitor the excited state population as a function of time. When, C comes into contact with W (within Van der Waals distance) due to stochastic collisions, an excited-state electron is transferred from the triplet state of W to C. As explained in Supporting Methods S2, and in Sizemore et al. 201536, the measured relaxation time is related to the intra-molecular contact formation time between W and C in the protein. The contribution to the triplet state relaxation times due to C-W quenching via electron transfer, are reported in Fig. 1D. These were obtained by fitting W triplet state relaxation curves as illustrated in Supporting Figs. S1 and S2, and taking into account the natural lifetime of the triplet state in the absence of C, measured on single W variants (see Supporting Methods S2).

Each value is the result of globally fitting multiple, repeated measurements, under given solution conditions. All raw data and their fits are shown in Supporting Figs. S3 through S16. The corresponding fitting parameters, obtained for each variant under each solution condition, are reported in Supporting Table S2. Solution conditions used in PET experiments of Fig. 1 were, for reference conditions: 15 mM Tris, 150 mM NaCl, 1 mM TCEP pH 7.6; for low pH measurements: 20 mM NaAc, 150 mM NaCl, 1 mM TCEP pH 4.0; for high salt concentrations: 15 mM Tris, 500 mM NaCl, 1mM TCEP pH 7.6.

Circular Dichroism

Circular Dichroism (CD) spectra reported in Fig. 1D inset were measured as described in Supporting Methods S3. To improve signal to noise, and access low wavelengths which are particularly important for the interpretation of IDP spectra, at each pH we carried out measurements for a series of protein concentrations and combined them. Measurements at high salt concentration were not possible due to the high absorbance of these samples below 210nm, a range which is most important for the interpretation of IDP spectra. Solution conditions used for spectra of Fig. 1 were: 10mM NaPO4, 150 mM, NaF for pH 7.6 measurements, and 20 mM NaAc, 150 mM NaF for pH 4.0 measurements.

Coarse-grained molecular dynamics

We applied a coarse-grained model based on the original HPS model43 to NTAIL. Each amino acid is represented by a bead with charge (+1, 0, −1) and hydropathy. There are three types of interactions in the HPS model: bonded interactions, electrostatic interactions, short-range pairwise interactions. The electrostatic interactions are modeled using a Coulombic term with Debye-Hückel43 electrostatic screening to account for salt concentration. The short-range pairwise potential accounts for both protein-protein and protein-solvent interactions, which was optimized using the experimental radius of gyration of NTAIL6. We further added additional terms for angle and dihedral preference so that the secondary structure of the MoRE region can be captured (see Supporting Methods S5). The HOOMD-Blue software v2.9.365 together with the azplugins (https://github.com/mphowardlab/azplugins) were used for running the molecular dynamics simulations. All simulations were run using a Langevin thermostat with a friction coefficient of 0.01 ps−1, a time step of 10 fs and a temperature of 298 K for 5 μs. The first 0.5 μs were dumped for equilibration.

All atom molecular dynamics

We performed all molecular dynamics (MD) simulations on three full length NTAIL variants. The wild type (WT) NTAIL (401–525) simulations were performed including an N-terminal hexa-histidine (His6) tag, while the latter was not included in the case of variants C3W2 (i.e. S482C-Y518W) and C4W2 (i.e. S488C-Y518W) to match the available experimental data as much as possible. All simulations were performed using the GROMACS 2019 simulation package66. Each NTAIL variant was simulated using two different force fields (CHARMM36m51 and AMBER99SB-disp49) and at least three replicas per force field started from different initial conformations. The WT NTAIL simulations were performed only with the CHARMM36m force field. For the Amber force field validation, we used previously published wild type NTAIL simulation trajectories26. The simulations of the two variants in Amber was carried out using force field parameters adapted to the GROMACS simulation package. Both the WT trajectories and the force field parameters were kindly provided by Piana et al. Each simulation was carried out using periodic boundary conditions in a box filled with explicit solvent molecules consisting of either optimal point charge (OPC55, for CHARMM36m) or a modified four point-charge (TIP4P50 for AMBER99SB-disp) water models as well as Na+ and Cl− ions corresponding to an ion concentration of 150 mM. The total size of simulated systems was approximately 120,000 to 200,000 atoms. Total simulation time was more than 40 μs per variant with conformations recorded after every nanosecond. Hydrogen bond vibrations were constrained using virtual atom sites to enable a 4 fs time step during simulations. All simulations were kept at 1 atm pressure and 298 K temperature. Further details on the MD simulation parameters and system preparation are provided in the Supporting Methods S7 and S9. Analysis methods related to the comparison to experimental data, calculation of mutual information, contact maps, and free-energy landscapes are described in the Supporting Methods S11, S1921.

Supplementary Material

Supplement 1
media-1.pdf (8.2MB, pdf)

Acknowledgements

L.O, J.K., G.K., S.L. and S.M.V. acknowledge the support from the National Institutes of Health (R01GM120537). A.C.V., H.G., L.V.B. and G.N. were supported by the Max Planck Society and by the German Science Foundation (DFG), Excellence Strategy Grant MBExC 2067/1; computer time was provided by the Max Planck Computing and Data Facility. G.N. was additionally supported by the Alexander von Humboldt-Foundation. A.C.V. was additionally supported by the European Union’s Horizon 2020 Framework Programme for Research and Innovation, Specific Grant Agreement No. 945539 (Human Brain Project SGA3). W.Z. acknowledges the support from the National Institutes of Health (R35GM146814) and the research computing at Arizona State University.

Data availability.

The implementation of the coarse-grained HPS model can be downloaded from https://github.com/wzhenglab/ntail. Other data supporting the findings of this study are available from the corresponding authors upon reasonable request.

References.

  • 1.Gutsche I. et al. Near-atomic cryo-EM structure of the helical measles virus nucleocapsid. Science (1979) 348, 704 (2015). [DOI] [PubMed] [Google Scholar]
  • 2.Bloyet L. M. The nucleocapsid of paramyxoviruses: Structure and function of an encapsidated template. Viruses 13, 2465 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Longhi S., Bloyet L. M., Gianni S. & Gerlier D. How order and disorder within paramyxoviral nucleoproteins and phosphoproteins orchestrate the molecular interplay of transcription and replication. Cellular and Molecular Life Sciences 74, 3091 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Karlin D., Ferron F., Canard B. & Longhi S. Structural disorder and modular organization in Paramyxovirinae N and P. Journal of General Virology 84, 3239 (2003). [DOI] [PubMed] [Google Scholar]
  • 5.Kingston R. L., Baase W. A. & Gay L. S. Characterization of Nucleocapsid Binding by the Measles Virus and Mumps Virus Phosphoproteins. J Virol 78, 8630 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Longhi S. et al. The C-terminal domain of the measles virus nucleoprotein is intrinsically disordered and folds upon binding to the C-terminal moiety of the phosphoprotein. Journal of Biological Chemistry 278, 18638–18648 (2003). [DOI] [PubMed] [Google Scholar]
  • 7.Jensen M. R. et al. Intrinsic disorder in measles virus nucleocapsids. Proc Natl Acad Sci U S A 108, 9839–9844 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Johansson K. et al. Crystal Structure of the Measles Virus Phosphoprotein Domain Responsible for the Induced Folding of the C-terminal Domain of the Nucleoprotein. Journal of Biological Chemistry 278, 44567 (2003). [DOI] [PubMed] [Google Scholar]
  • 9.Kingston R. L., Hamel D. J., Gay L. S., Dahlquist F. W. & Matthews B. W. Structural basis for the attachment of a paramyxoviral polymerase to its template. Proc Natl Acad Sci U S A 101, 8301–8306 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Sourimant J., Thakkar V. D., Cox R. M. & Plemper R. K. Viral evolution identifies a regulatory interface between paramyxovirus polymerase complex and nucleocapsid that controls replication dynamics. Sci Adv 6, eaaz1590 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Guseva S., Milles S., Blackledge M. & Ruigrok R. W. H. The nucleoprotein and phosphoprotein of measles virus. Front Microbiol 10, 1832 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Gondelaud F. et al. Functional benefit of structural disorder for the replication of measles, Nipah and Hendra viruses. Essays Biochem 66, 915 (2022). [DOI] [PubMed] [Google Scholar]
  • 13.Bloyet L. M. et al. Modulation of Re-initiation of Measles Virus Transcription at Intergenic Regions by PXD to NTAIL Binding Strength. PLoS Pathog 12, e1006058 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Cox R. & Plemper R. K. The paramyxovirus polymerase complex as a target for next-generation anti-paramyxovirus therapeutics. Frontiers in Microbiology vol. 6 459 Preprint at 10.3389/fmicb.2015.00459 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Huang W., Ravikumar K. M., Chance M. R. & Yang S. Quantitative mapping of protein structure by hydroxyl radical footprinting-mediated structural mass spectrometry: A protection factor analysis. Biophys J 108, 107–115 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Gely S. et al. Solution structure of the C-terminal X domain of the measles virus phosphoprotein and interaction with the intrinsically disordered C-terminal domain of the nucleoprotein. Journal of Molecular Recognition 23, 435 (2010). [DOI] [PubMed] [Google Scholar]
  • 17.Belle V. et al. Mapping α-helical induced folding within the intrinsically disordered C-terminal domain of the measles virus nucleoprotein by site-directed spin-labeling EPR spectroscopy. Proteins: Structure, Function and Genetics 73, 973 (2008). [DOI] [PubMed] [Google Scholar]
  • 18.Kavalenka A. et al. Conformational analysis of the partially disordered measles virus N TAIL-XD complex by SDSL EPR spectroscopy. Biophys J 98, 1055 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Wang Y. et al. Multiscaled exploration of coupled folding and binding of an intrinsically disordered molecular recognition element in measles èirus nucleoprotein. Proc Natl Acad Sci U S A 110, E3743 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Dosnon M. et al. Demonstration of a folding after binding mechanism in the recognition between the measles virus NTAIL and X domains. ACS Chem Biol 10, 795 (2015). [DOI] [PubMed] [Google Scholar]
  • 21.Bourhis J.-M. et al. The intrinsically disordered C-terminal domain of the measles virus nucleoprotein interacts with the C-terminal domain of the phosphoprotein via two distinct sites and remains predominantly unfolded. Protein Science 14, 1975 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Gruet A. et al. Fuzzy regions in an intrinsically disordered protein impair protein-protein interactions. FEBS Journal 283, 576–594 (2016). [DOI] [PubMed] [Google Scholar]
  • 23.Bignon C., Troilo F., Gianni S. & Longhi S. Modulation of measles virus NTAIL interactions through fuzziness and sequence features of disordered binding sites. Biomolecules 9, 8 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Troilo F., Bonetti D., Bignon C., Longhi S. & Gianni S. Understanding Intramolecular Crosstalk in an Intrinsically Disordered Protein. ACS Chem Biol 14, 337 (2019). [DOI] [PubMed] [Google Scholar]
  • 25.Mukhopadhyay R. & Hoh J. H. AFM force measurements on microtubule-associated proteins: The projection domain exerts a long-range repulsive force. FEBS Lett 505, 374 (2001). [DOI] [PubMed] [Google Scholar]
  • 26.Robustelli P., Piana S. & Shaw D. E. Mechanism of Coupled Folding-upon-Binding of an Intrinsically Disordered Protein. J Am Chem Soc 142, 11092–11101 (2020). [DOI] [PubMed] [Google Scholar]
  • 27.Gruet A. et al. Dissecting partner recognition by an intrinsically disordered protein using descriptive random mutagenesis. J Mol Biol 425, 3495 (2013). [DOI] [PubMed] [Google Scholar]
  • 28.Cai L. & Zhou H. X. Theory and simulation on the kinetics of protein-ligand binding coupled to conformational change. Journal of Chemical Physics 134, 105101 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.White J. T., Li J., Grasso E., Wrabl J. O. & Hilser V. J. Ensemble allosteric model: Energetic frustration within the intrinsically disordered glucocorticoid receptor. Philosophical Transactions of the Royal Society B: Biological Sciences 373, 20170175 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Berlow R. B., Dyson H. J. & Wright P. E. Hypersensitive termination of the hypoxic response by a disordered protein switch. Nature 543, 447 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Chowdhury A., Nettels D. & Schuler B. Interaction Dynamics of Intrinsically Disordered Proteins from Single-Molecule Spectroscopy. Annu Rev Biophys 52, 433–462 (2023). [DOI] [PubMed] [Google Scholar]
  • 32.Staby L. et al. Flanking Disorder of the Folded αα-Hub Domain from Radical Induced Cell Death1 Affects Transcription Factor Binding by Ensemble Redistribution: Flanking disorder tunes the RCD1-RST ensemble. J Mol Biol 433, 167320 (2021). [DOI] [PubMed] [Google Scholar]
  • 33.Peng Y. et al. A Metastable Contact and Structural Disorder in the Estrogen Receptor Transactivation Domain. Structure 27, 229–240 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Buscaglia M., Schuler B., Lapidus L. J., Eaton W. A. & Hofrichter J. Kinetics of intramolecular contact formation in a denatured protein. J Mol Biol 332, (2003). [DOI] [PubMed] [Google Scholar]
  • 35.Vaiana S., Best R. B., Yau W.-M., Eaton W. A. & Hofrichter J. Evidence for a partially structured state of the amylin monomer. Biophys. J. 97, 2948–2957 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Sizemore S. M., Cope S. M., Roy A., Ghirlanda G. & Vaiana S. M. Slow Internal Dynamics and Charge Expansion in the Disordered Protein CGRP: A Comparison with Amylin. Biophys J 109, 1038 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Lapidus L. J., Eaton W. A. & Hofrichter J. Measuring the rate of intramolecular contact formation in polypeptides. Proc. Natl. Acad. Sci. U.S.A. 97, 7220–7225 (2000). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Schuler B. & Hofmann H. Single-molecule spectroscopy of protein folding dynamics—expanding scope and timescales. Curr. Opin. Struct. Biol. 23, 36–47 (2013). [DOI] [PubMed] [Google Scholar]
  • 39.Flory P. J. Principles of Polymer Chemistry. Principles of Polymer Chemistry (Cornell University Press, 1953). [Google Scholar]
  • 40.Zheng W. et al. Inferring properties of disordered chains from FRET transfer efficiencies. Journal of Chemical Physics 148, 123329 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Sawle L. & Ghosh K. A theoretical method to compute sequence dependent configurational properties in charged polymers and proteins. J. Chem. Phys. 143, 85101 (2015). [DOI] [PubMed] [Google Scholar]
  • 42.Das R. K. & Pappu R. V. Conformations of intrinsically disordered proteins are influenced by linear sequence distributions of oppositely charged residues. Proc. Natl. Acad. Sci. U.S.A. 110, 13392–13397 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Dignon G. L., Zheng W., Kim Y. C., Best R. B. & Mittal J. Sequence determinants of protein phase behavior from a coarse-grained model. PLoS Comput. Biol. 14, e1005941 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Debye P. & Hückel E. De la theorie des electrolytes. I. abaissement du point de congelation et phenomenes associes. Physikalische Zeitschrift 24, 185–206 (1923). [Google Scholar]
  • 45.Nagy G., Igaev M., Jones N. C., Hoffmann S. V. & Grubmüller H. SESCA: Predicting Circular Dichroism Spectra from Protein Molecular Structures. J Chem Theory Comput 15, 5087 (2019). [DOI] [PubMed] [Google Scholar]
  • 46.Fossat M. J. & Pappu R. V. Q-Canonical monte carlo sampling for modeling the linkage between charge regulation and conformational equilibria of peptides. Journal of Physical Chemistry B 123, 6952 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Zheng W. et al. Molecular Details of Protein Condensates Probed by Microsecond Long Atomistic Simulations. Journal of Physical Chemistry B 124, 11671 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Best R. B., Zheng W. & Mittal J. Balanced protein-water interactions improve properties of disordered proteins and non-specific protein association. J. Chem. Theor. Comput. 10, 5113–5124 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Robustelli P., Piana S. & Shaw D. E. Developing a molecular dynamics force field for both folded and disordered protein states. Proc. Natl. Acad. Sci. U.S.A. 115, E4758–E4766 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Piana S., Donchev A. G., Robustelli P. & Shaw D. E. Water dispersion interactions strongly influence simulated structural properties of disordered protein states. J. Phys. Chem. B 119, 5113–5123 (2015). [DOI] [PubMed] [Google Scholar]
  • 51.Huang J. et al. CHARMM36m: an improved force field for folded and intrinsically disordered proteins. Nat Methods 14, 71 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Henriques J. & Skepö M. Molecular dynamics simulations of intrinsically disordered proteins: On the accuracy of the TIP4P-D water model and the representativeness of protein disorder models. J Chem Theory Comput 12, 3407–3415 (2016). [DOI] [PubMed] [Google Scholar]
  • 53.Svergun D., Barberato C. & Koch M. H. J. CRYSOL - a program to evaluate X-ray solution scattering of biological molecules from atomic coordinates. J. Appl. Cryst. 28, 768–773 (1995). [Google Scholar]
  • 54.Shen Y. & Bax A. SPARTA+: a modest improvement in empirical NMR chemical shift prediction by means of an artificial neural network. J. Biomol. NMR 48, 13–22 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Shabane P. S., Izadi S. & Onufriev A. V. General Purpose Water Model Can Improve Atomistic Simulations of Intrinsically Disordered Proteins. J Chem Theory Comput 15, 2620 (2019). [DOI] [PubMed] [Google Scholar]
  • 56.Hopf T. A. et al. The EVcouplings Python framework for coevolutionary sequence analysis. Bioinformatics 35, 1582–1584 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Bignon C., Troilo F., Gianni S. & Longhi S. Partner-Mediated Polymorphism of an Intrinsically Disordered Protein. J Mol Biol 430, 2493–2507 (2018). [DOI] [PubMed] [Google Scholar]
  • 58.Milles S. et al. An ultraweak interaction in the intrinsically disordered replication machinery is essential for measles virus function. Sci Adv 4, eaat7778 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Krumm S. A., Takeda M. & Plemper R. K. The measles virus nucleocapsid protein tail domain is dispensable for viral polymerase recruitment and activity. Journal of Biological Chemistry 288, 29943 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Guseva S. et al. Measles virus nucleo- and phosphoproteins form liquid-like phase-separated compartments that promote nucleocapsid assembly. Sci Adv 6, eaaz7095 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Zhou Y., Su J. M., Samuel C. E. & Ma D. Measles Virus Forms Inclusion Bodies with Properties of Liquid Organelles. J Virol 93, e00948–19 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Kolakofsky D., Le Mercier P., Iseni F. & Garcin D. Viral RNA polymerase scanning and the gymnastics of Sendai virus RNA synthesis. Virology 318, 463 (2004). [DOI] [PubMed] [Google Scholar]
  • 63.Brunel J. et al. Sequence of Events in Measles Virus Replication: Role of Phosphoprotein-Nucleocapsid Interactions. J Virol 88, 10851 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Bugge K. et al. Interactions by Disorder – A Matter of Context. Front Mol Biosci 7, 110 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Anderson J. A., Glaser J. & Glotzer S. C. HOOMD-blue: A Python package for high-performance molecular dynamics and hard particle Monte Carlo simulations. Comput Mater Sci 173, 109363 (2020). [Google Scholar]
  • 66.Abraham M. J. et al. Gromacs: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 1, 19 (2015). [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplement 1
media-1.pdf (8.2MB, pdf)

Data Availability Statement

The implementation of the coarse-grained HPS model can be downloaded from https://github.com/wzhenglab/ntail. Other data supporting the findings of this study are available from the corresponding authors upon reasonable request.


Articles from bioRxiv are provided here courtesy of Cold Spring Harbor Laboratory Preprints

RESOURCES