Abstract
Cooperative interactions play an important role in recognition and binding in macromolecular systems. In this study, we find that cross-correlated fluctuations can be used to identify cooperative networks in a protein-RNA system. The dynamics of the RRM-containing protein U1A–stem loop 2 RNA complex have been calculated theoretically from a 10 ns molecular dynamics (MD) simulation. The simulation was analyzed by calculating the covariance matrix of all atomic fluctuations. These matrix elements are then presented in the form of a two-dimensional grid, which displays fluctuations on a per-residue basis. The results indicate the presence of strong, selective cross-correlated fluctuations throughout the RRM in U1A-RNA, which correspond well with the results from a statistical covariance analysis of 330 aligned RRM sequences and previous biophysical studies in which a multiplicity of cooperative networks have been reported. Moreover, the MD results indicate that the various networks identified in separate individual experiments are fluctuationally correlated into a hyper-network encompassing most of the RRM. This method has significant implications as a predictive tool regarding cooperativity in the protein-nucleic acid recognition process.
Keywords: cooperativity, cooperative network, cross-correlation, collective atomic fluctuations, U1A, protein-RNA recognition, molecular dynamics
1. Introduction
Protein-RNA complexes are involved in most steps of gene expression, including the editing, transportation, translation, modification and degradation of RNA (Varani and Nagai, 1998; Chen and Varani, 2005). As such, insight into the protein-RNA recognition process is necessary for a complete understanding of gene expression. However, the high affinity and specificity with which proteins are able to recognize RNAs is a complicated issue to probe. Factors that contribute to protein-RNA recognition include the structures, thermodynamics, cooperative effects and dynamical motions of the macromolecules involved, all of which are highly interdependent. To probe this issue, we investigate the U1A-RNA complex by calculating the cross-correlations of atomic fluctuations derived from the analysis of a molecular dynamics (MD) simulation. We then suggest a link between the fluctuational cross-correlations and cooperative networks of interactions between residues in the U1A-RNA complex. This method enables one to link functional dynamics to cooperativity in the pursuit of a better understanding of protein-RNA recognition.
2. Background
The U1A-RNA complex has been well-characterized by numerous structural and chemical studies, making it a paradigm for the study of protein-RNA recognition, interaction and binding. The protein U1A recognizes its cognate RNA through one of the most common eukaryotic protein domains, the RNA recognition motif (RRM), also known as the RNA binding domain (RBD) or the ribonucleoprotein (RNP) domain. The general structure of the RRM is an antiparallel β-sheet flanked by two α-helices forming a βαββαβ motif (Maris et al., 2005).
U1A binds stem loop 2 (SL2) of U1 snRNA with high affinity and specificity (KD < 10−9 M) (Green, 1991; Tsai et al., 1991; Hall, 1994; Stark et al., 2001). The complex of the N-terminal RRM of U1A with SL2 of U1 snRNA (hereafter referred to as U1A-RNA) has been solved by X-ray crystallography (Oubridge et al., 1994). The primary sequence and secondary structure of U1A are presented in Fig. 1. The tertiary structure of the U1A-RNA complex and the secondary structure of SL2 RNA are shown in Fig. 2. Upon complex formation, two highly conserved aromatic amino acids on the β-sheet stack with bases in the single-stranded loop region of the RNA: Tyr13 stacks with C5 and Phe56 stacks with A6, which stacks with C7 (Oubridge et al., 1994). A C-terminal α-helix, labeled α C in Fig. 2, is not part of the RRM, but has been shown to be important for the binding of RNA by U1A (Avis et al., 1996; Zeng and Hall, 1997; Kranz and Hall, 1998; Mittermaier et al., 1999). Loop 3 protrudes through the SL2 RNA loop upon binding (Oubridge et al., 1994) and loop 1, loop 3 and residues Thr89-Asp90-Ser91 (‘the TDS linker’) have been shown to be important for specificity in the binding of SL2 RNA by U1A (Scherly et al., 1990; Bentley and Keene, 1991; Boelens et al., 1991; Jessen et al., 1991; Laird-Offringa and Belasco, 1995; Allain et al., 1997). We have previously reported both experimental and molecular dynamics (MD) simulations to probe structural adaptations and the energetic origins of affinity and specificity in the U1A-stem loop 2 RNA complex (Nolan et al., 1999; Luchansky et al., 2000; Blakaj et al., 2001; Pitici et al., 2002; Shiels et al., 2002; Tuite et al., 2002; Gayle and Baranger, 2002; Zhao and Baranger, 2003; Zhao et al.).
Recent studies have indicated that U1A-RNA recognition and binding occur not only as a result of direct interactions, but are also due to long-range effects that are not identifiable merely by inspection of the tertiary structure. One such effect is cooperativity, which plays a functional role in a number of macromolecular activities, including recognition and binding (Di Cera, 1998b). First introduced by Wyman (Wyman, 1948; Wyman and Gill, 1990), the concept of cooperativity represents an interdependence of two or more different groups in a system causing a higher binding affinity than would be expected from summing the association energies of the individual parts (nonadditivity). Common ways to determine cooperativity in a macromolecular system are through the use of site directed mutagenesis, alanine scanning, and double mutant cycles to determine the presence of energetic coupling between residues of interest (Ackers and Smith, 1985; Horovitz and Fersht, 1990; Di Cera, 1998a). A computational theory aimed at the elucidation of cooperative effects in proteins has been set forth in an extensive series of papers by Freire and coworkers (Murphy and Freire, 1992; Hilser et al., 1998; Freire, 1999; Luque et al., 2002). In these calculations, an ensemble of partially unfolded structures is generated using a crystal or NMR structure as a template. Configurational energies and related quantities of the structures are calculated from empirical energy functions fitted to solvent accessible surface areas. The configurational energies obtained in this manner are pairwise additive, and cooperative effects are elucidated by sensitivity analysis on a per residue basis. This approach goes directly from an enumeration of partially folded states to thermodynamic properties without explicitly generating structures, whereas the study described in this article is structure-based.
A series of recent studies have provided biophysical evidence for three cooperative networks of interactions throughout U1A that play a role in binding and specificity (Fig. 5). The first network was hypothesized by Kranz and Hall (1999) and involves interactions from Tyr13 through the RNA to the C-term residues (Fig. 5A). This hypothesis is based on a study in which double mutant cycles were used to reveal energetic coupling between Tyr13 and the C-terminal tail of U1A when bound to RNA (Kranz and Hall, 1998). The contribution of Phe56 has also been added to the cooperative network detailed in Fig. 5A. This is justified based on several studies. Kranz and Hall (1999) found energetic coupling found between Tyr13 and Phe56 based on double mutant cycles. In addition, studies by Baranger and coworkers showed that mutation of Phe56 to Ala altered the U1A-RNA binding affinity by more than would be anticipated from a loss in stacking interactions (Nolan et al., 1999; Tuite et al., 2002) and that A6 of RNA is energetically coupled to Phe56 (Shiels et al., 2002). The second network, also hypothesized by Kranz and Hall (1999), involves interactions from Tyr13 through Gln54 and the loop 3 residues through the RNA (Fig. 5B). This hypothesis is based on a study in which double mutant cycles were used to show energetic coupling between Tyr13 and Gln54, and NMR relaxation experiments showed that mutations in Tyr13, Gln54 and Phe56 affected the dynamics of loop 3 (Kranz and Hall, 1999). Laird-Offringa and coworkers have also performed studies reinforcing the role of Gln54 in the U1A-RNA cooperative networks (Katsamba et al., 2002; Law et al., 2005). The third cooperative network in U1A-RNA involves interactions from Gly53 through loop 1 and loop 3 to the TDS linker (Fig. 5C). This network is hypothesized due to a study in which Showalter and Hall (2002) used NMR, MD and reorientational eigenmode dynamics (RED) techniques to show that the motions of loop 1, loop 3 and loop 6 residues (the TDS linker) are correlated and that mutations in Gly53 affected those correlations. Using the same techniques, they showed that Gly53 has a specific role in RNA binding and that mutations of this residue weaken RNA binding, indicating decoupling of intradomain motions (Showalter and Hall, 2004). While this work was in progress, Showalter and Hall (2005) used MD and RED to identify correlated motions in the U1A-RNA complex and suggested the motions are a reflection of an extensive cooperative network of interactions in this system.
Generalizing the role of networks in protein-RNA binding, Crowder et al. (2001) performed a statistical covariance analysis (Lockless and Ranganathan, 1999) of 330 RRMs (including U1A) to probe the RNA-binding specificity of proteins that contain these binding motifs. They identified pairs of positions that are functionally coupled, revealing a network of covariant residue pairs beyond those determined in U1A cooperativity studies (vide supra). The Crowder et al. study implicates residues not only on the RNA-binding surface of the RRMs, but also those in the hydrophobic cores of the proteins, indicating extensive networks between residues in the interior of the protein and those in contact with RNA. These results expand upon the important role of functionally linked networks of residues in the binding of RNA by RRMs and specify positions present throughout 330 RRMs that are particularly important for this process.
In this study, we pursue the relationship of fluctuational cross-correlations and cooperativity. Functional dynamics in macromolecular systems are not easy to probe experimentally, and are impossible to determine from a visual inspection of crystal or NMR structures. Thus, computational theory and methodologies have a unique vantage point on this problem. MD simulations serve as a link between structure and dynamics by providing detailed atomic motions as a function of time (Karplus and Petsko, 1990; Doniach and Eastman, 1999; Karplus and McCammon, 2002; Karplus and Kuriyan, 2005). The extent to which atomic fluctuations are correlated with one another can then be extracted from an MD simulation by calculation of the covariance matrix of all atomic fluctuations (McCammon and Harvey, 1986; Brooks III et al., 1988). The matrix elements may then be presented graphically in a dynamical cross-correlation map (DCCM, perhaps better described as a fluctuational cross-correlation map), giving a picture of the correlated fluctuations throughout the course of the MD trajectory. The first studies utilizing this technique involved analysis of correlated motions in cytochrome c (McCammon, 1984), ribonuclease A with two different substrates (Brünger et al., 1985), HIV-1 protease (Harte et al., 1990; Harte et al., 1992; Swaminathan et al., 1991), and BPTI (Ichiye and Karplus, 1991). Since these pioneering studies, cross-correlations and DCCMs have been used to analyze the fluctuations of many diverse systems, including proteins (Arcangeli et al., 1999; Parchment and Essex, 2000; Arcangeli et al., 2001; Rizzuti et al., 2001; Young et al., 2001; Luo and Bruice, 2002; Zoete et al., 2002), protein-nucleic acid complexes (Suenaga et al., 2000; Trylska et al., 2005), and enzymes (Bahar et al., 1997; Radkiewicz and Brooks III, 2000; Rod et al., 2003; Sulpizi et al., 2003; Agarwal, 2004; Gunasekaran and Nussinov, 2004; Schiøtt, 2004; Wong et al., 2004; Fuxreiter et al., 2005; Gorfe and Caflisch, 2005; Ma et al., 2005). A recent study analyzing both the B1 domain of protein G and ubiquitin suggests the correlated motions obtained from MD simulations agree well with measured NMR data (Lange et al., 2005).
3. Methods and Calculations
3.1. MD Methods
The starting structure for the MD simulation of U1A-RNA was based on the x-ray cocrystal structure of the N-terminal RRM of U1A bound to SL2 of U1 snRNA solved at 1.92 Å resolution (Oubridge et al., 1994), PDB ID (Berman et al., 2000): 1URN. Biological unit 2 was chosen for the initial structure as it contains the most complete structural information for SL2 RNA. The U1A protein was extended from the crystal structure construct to obtain a structure containing residues 2 – 102, and two point mutations (H31Y and R36Q) were introduced to revert the protein to the wild type sequence. Details of these procedures are described in Pitici et al. (2002).
Trajectories were computed using the AMBER (Pearlman et al., 1995; Case et al., 2004) suite of programs with the parm96 (Kollman et al., 1997) force field. The U1A-RNA system was solvated in a box of explicit TIP3P (Jorgensen et al., 1983) water molecules that extended a minimum distance of 12 Å from the solute atoms. 13 Na+ ions were initially added to neutralize the system. A salt concentration of 250 mM was then obtained by adding 45 Na+ ions and 45 Cl− ions to the system. The ions were randomized around the solute using the randomizeions command such that the ions could be no closer than 5 Å to the solute and no closer than 3 Å to each other. Minimization and MD simulation of the U1A-RNA complex followed the protocol described by Pitici et al. (2002). The results presented here are based on a production run of 10 ns.
3.2. Cross-Correlation Computation and DCCM Details
The extent to which the fluctuations of a system are correlated is dependent on the magnitude of the cross-correlation coefficient (Brooks III et al., 1988; McCammon and Harvey, 1986), Cij, given by Eq. 1:
(1) |
where i and j may be any two atoms, residues or domains, Δri and Δrj are displacement vectors of i and j, and the angle brackets denote an ensemble average. This involves calculation of all inter-atomic cross-correlations of atomic fluctuations, forming a matrix whose elements Cij may be displayed in a graphical representation as a dynamical (fluctuational) cross-correlation map, or DCCM. If Cij = 1 the fluctuations of i and j are completely correlated (same period and same phase), if Cij = −1 the fluctuations of i and j are completely anticorrelated (same period and opposite phase), and if Cij = 0 the fluctuations of i and j are not correlated.
In the DCCMs reported in this paper, the average equilibrated structure from MD has been used as the reference. Only cross-correlation coefficients greater than 0.25 and less than −0.25 are plotted, with black indicating Cij values of ±0.75 – 1.00, slate gray indicating Cij values of ±0.50 – 0.75 and light gray indicating Cij values of ±0.25 – 0.50. Typical characteristics of DCCMs include a line of strong cross-correlations along the diagonal, cross-correlations emanating from the diagonal, and off-diagonal cross-correlations. The strong correlated fluctuations along the diagonal occur for i = j, where Cij is always equal to 1.00. Positive correlations emanating from the diagonal indicate fluctuational correlations between contiguous residues, typically within a secondary structure domain. Signature secondary structure patterns include a triangular pattern for an α-helix and a plume for a β-strand. Off-diagonal positive and negative correlations indicate fluctuational correlations between domains of non-contiguous residues. These correlations provide dynamical information about the structure that may not necessarily be apparent from inspection of an average or experimentally derived tertiary structure.
3.3. Technical Details
The choice of fitting procedure used to remove overall rotation from the coordinate sets derived from a MD trajectory has been a point of discussion in the literature. Karplus and coworkers (Ichiye and Karplus, 1991; Karplus and Ichiye, 1996; Zhou et al., 2000) have shown that minimization of the root mean square deviation (RMSD) of all Cα is the most appropriate choice to obtain a complete picture of the fluctuational correlations present within the system. Comparison of their results to those from normal mode analysis (Ichiye and Karplus, 1991) as well as an additional study using MD at zero-angular momentum (Zhou et al., 2000) revealed no spurious correlations when the structures are superimposed over all Cα. Prompers and Brüschweiler (2002) used an isotropically distributed ensemble analysis method which separates internal motion from overall motion and obtained results similar to those obtained by superposition over all Cα atoms. However, Hünenberger et al. (1995) suggest superposition of structures over the least mobile (secondary structure) Cα atoms is a more appropriate choice for fitting as it eliminates ‘artifactual’ correlations observed at long distances in the system. Schieborr and Rüterjans (2001) observe a similar result by comparing to a bias-free method of removing internal and overall motions. Abseher and Nilges (1998) showed that superposition of structures over the least mobile Cα atoms gives results more similar to those obtained from distance space results. Additionally, they showed that superposition over one rigid part of a molecule (for example, one of the monomers in a dimer) causes the misleading appearance of increased motional correlations in the other portions of the molecule. With this knowledge, the cross-correlation coefficients for the U1A-RNA complex were computed after superposition of structures over all Cα atoms and the resultant DCCM was compared to the DCCM computed after superposition over the Cα atoms of only the βαββαβ RRM residues. The resultant DCCMs were virtually identical (data not shown). As such, the results presented herein are those derived from superposition over all Cα atoms to avoid overemphasizing the motion of other structural elements (e.g., helix C).
Additional discussion is warranted regarding several other issues that must be considered when computing the cross-correlation coefficients for a molecule. One issue involves the choice of atoms for this type of analysis. An all-atom cross-correlation matrix is significantly memory- and data-intensive for any system on the scale of a macromolecule, so analysis is typically done by grouping atoms. Grouping for a protein may be done all-atom by residue, but is typically done for backbone atoms by residue. Sensitivity of this analysis to the choice of atom grouping for the protein was tested using the U1A protein. Grouping choices included all atoms by residue, backbone N-Cα-C-O atoms by residue, backbone N-Cα-C atoms by residue, and backbone Cα atoms by residue. All of the DCCMs that involved only backbone atoms looked very similar (data not shown). The cross-correlation coefficients for all atoms by residue showed the same patterns seen for the others, however the patterns were rougher and less correlated due to the increase in motion from inclusion of the more dynamic side chains in the analysis. As such, the DCCMs presented in this study have been grouped using the backbone N-Cα-C atoms by residue for U1A. Sensitivity of this analysis to the choice of atom grouping for the RNA was tested using SL2 RNA. Grouping choices included all-atom by nucleotide and the sugar-phosphate backbone P-O5′-C5′-C4′-C3′-O3′ atoms. Resultant cross-correlation coefficients for these two different grouping choices were similar (data not shown). For the DCCMs presented in this study, grouping for SL2 RNA has been done all-atom by nucleotide because the base-paired stem portion of the RNA revealed stronger correlations than for the results computed over the sugar-phosphate backbone.
The time scale over which correlation data are collected is another important issue to consider when computing cross-correlation coefficients. Fluctuational correlations will depend on the time scale over which the data is analyzed and different correlations may arise on different time scales. The DCCMs presented in this study are based on an ensemble average of one structure per picosecond compared to the equilibrated MD structure. This ensures that only the most important and/or consistent correlated fluctuations that are longer than 1 ps and shorter than 10 ns (or the length of the trajectory) will be observed in the results. As a test, block averages were computed for U1A over intervals of time for 5, 10, 25, 50, 100, 500 and 1000ps. The cross-correlation coefficients all looked very similar to those computed for the 1 ps time scale up to 100 ps (data not shown). At this point and for the larger time scales, the DCCMs become rougher and exhibit stronger correlations between much of the protein, a result of less sampling of the trajectory.
The final consideration of interest in this type of analysis is reproducibility of the results with other force fields. There have been several revision of the AMBER force field since this project was initiated (Wang et al., 2000; Okur et al., 2002). This issue was tested by running a simulation on U1A using the “Stony Brook” modification to the ff99 backbone torsions (Hornak et al., 2006) in AMBER (Case et al., 2006), f99SB, and comparing the cross-correlation coefficients computed from this analysis to those from the parm96 force field. A few residues exhibited slightly stronger correlated fluctuations in the DCCM from the ff99SB force field, but overall, the DCCMs were very similar. This indicates the dynamics of U1A are very similar and can be reproduced using these two different force fields.
3.4. Images
All molecular images were created using VMD (Humphrey et al., 1996) and rendered using POV-Ray (2004). All DCCM images were created using Molecular Dynamics (MD) Toolchest (Ravishanker et al., 1998) and WesGraph (Ehresman et al., 1995).
4. Results
4.1. MD results
A molecular dynamics trajectory was computed for the U1A-RNA complex for 10 ns. The root-mean-square deviation (RMSD) of the protein backbone and all RNA atoms from the equilibrated structure was calculated over the course of the trajectory and is plotted in Fig. 3. The RMSD stabilized after 0.6 ns and remained stable throughout the remainder of the simulation. Verification of the trajectory was provided by comparison to previous MD studies on the U1A-RNA system (Blakaj et al., 2001; Hermann and Westhof, 1999; Reyes and Kollman, 1999; Showalter and Hall, 2005; Tang and Nilsson, 1999). The average RMSD of 1.4 Å (computed for the protein backbone and over all RNA atoms) is consistent with the prior studies. The root-mean-square fluctuations (RMSF) over the course of the trajectory were computed on a residue or nucleotide basis with reference to the average structure and are plotted in Fig. 4. The largest fluctuations occur in the flexible N- and C-terminal regions of the U1A protein, and in the flexible stem and U8/C9/C10 nucleotides of SL2 RNA, nucleotides that do not interact with the protein. Additional analyses of this trajectory are currently in progress, including a complete analysis of the ion and hydration atmospheres.
4.2. DCCM of U1A-RNA
Cross-correlation coefficients were computed from the MD trajectory of the U1A-RNA complex using Eq. 1 and are presented in the DCCM shown in Fig. 6. For the DCCM results presented in this study, cross-correlation coefficients have been computed by residue over the center of mass of N, Cα and C backbone atoms in U1A and by nucleotide over the center of mass of all atoms in SL2 RNA, using the equilibrated structure as the reference structure. The analysis is based on an ensemble average of one structure per picosecond over the entire 10 ns trajectory. Structures for this analysis were superimposed by minimizing the RMSD over all Cα to the equilibrated structure. Positive correlations are collected in the upper triangle of the DCCM and negative correlations are collected in the lower triangle. Only correlations stronger than 0.25 are shown and darker colors indicate stronger correlated fluctuations where black indicates Cij values of ±0.75 – 1.00, slate gray indicates Cij values of ±0.50 – 0.75 and light gray indicates Cij values of ±0.25 – 0.50. Note that not only are protein-protein correlated fluctuations obtained, but protein-RNA and RNA-RNA correlations are obtained, as well. Red (positive correlated fluctuations) and blue (negative correlated fluctuations) boxes have been drawn on the plot in Fig. 6 to indicate locations of intramolecular U1A secondary structure intersections. Orange (positive correlated fluctuations) and green (negative correlated fluctuations) lines have been drawn on the plot to distinguish between SL2 stem and loop nucleotides. This DCCM provides a picture of the correlated fluctuations within the U1A-RNA system that occur throughout the 10 ns MD trajectory.
4.2.1. Protein – protein correlations
U1A intramolecular correlated fluctuations are shown in the large box in the lower left of Fig. 6 with residues labeled from 2 to 102. Signature triangle patterns emanating from diagonal are evident for the three α-helices in U1A; helix A includes residues Lys23 – Phe37, helix B includes Val62 – Met72, and helix C includes residues Asp90 – Lys98. The single pair of contiguous β-strands, β2 – β3, exhibits the signature plume of an antiparallel β-sheet emanating from the diagonal. The remainder of the β-strands comprising the β-sheet are seen in signature plumes that occur off-diagonal as they occur in non-contiguous residues; the direction of the plume depends on whether the β-strands in question run parallel or antiparallel. β1 includes residues Thr11 – Asn 15, β2 includes residues Ile40 – Val45, β3 includes residues Ala55 – Phe59, and β4 includes residues Arg83 – Tyr86. Note there is a significant amount of correlation off of the diagonal, indicating a large amount of fluctuational correlation between non-contiguous residues in the U1A protein. There is not a great deal of anticorrelated protein – protein fluctuation. The most significant is seen between the N- terminal residues with helix A, the C-terminal residues with loops 1 and 5, and the N- and C-terminal ends of loop 3 with β2 and β3.
4.2.2 Protein – RNA correlations
U1A positive and negative correlated fluctuations with SL2 RNA are shown in the rectangles at the top and on the right of Fig. 6, respectively, where U1A is labeled 2 to 102, the 5′ RNA stem is labeled −5 to −1, the single-stranded RNA loop is labeled 1 to 10, and the 3′ RNA stem is labeled 11 to 16. The fluctuations of the loop nucleotides of SL2 RNA are strongly correlated with those of the U1A β-sheet, especially β1 and β3, in which stacking interactions are taking place with the conserved aromatic residues: Tyr13/C5 and Phe56/A6/C7. Loop 1 reveals positive cross-correlations with nucleotides in the RNA loop preceding those involved in stacking with the protein, A1, U2, U3, and G4. Many of these correlations correspond with RNA-protein interactions observed in the X-ray crystal structure (Oubridge et al., 1994): U2 hydrogen bonds Glu19, U3 hydrogen bonds Asn16, and G4 hydrogen bonds Asn15 and Asn16, and interacts with Leu17 through a water-mediated hydrogen bond. However, the strongly correlated fluctuations between loop 1 and the RNA loop extend beyond that predicted by the contacts. Loop 3 reveals positive cross-correlations with the single-stranded nucleotides and the C-G base pair of the SL2 sequence that is recognized by U1A. As with loop 1, some of the U1A loop 3 – RNA cross-correlated fluctuations correspond to interactions shown in the crystal structure, while others extend beyond the contacts. Crystal structure contacts in this region include: Arg52 hydrogen bonds A1, G4 interacts with Leu49 through a water-mediated hydrogen bond and stacks with Gln54, and C5 hydrogen bonds Gln54. Single-stranded loop nucleotides also reveal some strong fluctuational correlations with loop 5 and helix C. Crystal contacts in this region include: C5 hydrogen bonds Lys88, A6 hydrogen bonds Thr89 and Ser90 and interacts with Lys88 through a water-mediated hydrogen bond, and C7 hydrogen bonds Ser91 and Asp92. Again, the positive fluctuational correlations in this region extend beyond the specific protein-RNA interactions. Loops 1 and 3 additionally have positive cross-correlations with stem. Crystal structure interactions in this region include: Lys20 and Lys22 interact with the sugar-phosphate backbone of the 5′ stem and Arg52 hydrogen bonds the closing C-G base pair. Again, cross-correlated fluctuations extend beyond the contacts. The fluctuations of the remainder of the stem are generally anticorrelated with those of the protein, except for loops 1 and 3.
4.2.3. RNA – RNA correlations
The SL2 RNA fluctuational cross-correlations with itself are shown in the small box in the top right of Fig. 6, and nucleotides are labeled from −5 to 16. The 5′ and 3′ stem nucleotides are strongly correlated, as is expected between Watson-Crick base-paired nucleotides. There are also positive cross-correlations seen within each 5′ and 3′ stem region which may be expected with intrastrand stacking. The single stranded loop nucleotides show anticorrelated fluctuations with stem, most noticeably with the dangling U16 nucleotide.
4.3. Correlated fluctuations and cooperativity
This analysis has demonstrated that using a DCCM to analyze a molecular dynamics trajectory can provide a useful picture of the collective atomic fluctuations taking place in the simulation. Next, we address the hypothesis that collective atomic fluctuations are indicative of cooperative effects by comparing the computed cross-correlation coefficients to observed cooperativity data in U1A-RNA. The issue of a cooperative network of interactions between residues in the U1A-SL2 RNA system that contributes to both binding and maintaining the complex has been primarily addressed by the work of Hall and coworkers (Kranz and Hall, 1998, 1999; Showalter and Hall, 2002, 2004; Showalter and Hall, 2005), Baranger and coworkers (Nolan et al., 1999; Shiels et al., 2002; Tuite et al., 2002; Zhao and Baranger, 2003), and Laird-Offringa and coworkers (Katsamba et al., 2002; Law et al., 2005). We have observed that the residues involved in strong fluctuational correlations in the U1A-RNA complex correspond well to those observed to be a part of the cooperative networks of interactions. (The cooperative interactions that contribute to binding and the extent to which these correlations change in U1A and RNA upon binding will be explored in a forthcoming study.)
The cooperative networks of interactions that emerge from the previous studies on U1A-RNA and the residues involved are summarized in Fig. 5. The first network (Fig. 5A) involves conserved residues Tyr13 and Phe56, through the RNA nucleotides C5, A6 and C7, to the C-terminal U1A residues, Lys96 – Phe101 (Kranz and Hall, 1998, 1999; Nolan et al., 1999; Shiels et al., 2002; Tuite et al., 2002; Zhao and Baranger, 2003; Law et al., 2005). The second network (Fig. 5B) involves Tyr13 to Gln54 to loop 3 to the RNA closing C-G base pair and A1 (Kranz and Hall, 1999; Katsamba et al., 2002; Law et al., 2005). The third network (Fig. 5C) involves Gly53, loop 1 and loop 3 through the RNA to Thr89-Asp90-Ser91 (the ‘TDS linker’) (Showalter and Hall, 2002, 2004).
The fluctuational correlations of the specific residues implicated in the cooperative networks are highlighted in the DCCM in Fig. 7A. The residues in the first network are outlined in red. Tyr13 is strongly correlated with C5, Phe56 is strongly correlated with A6, A6 is moderately correlated with C7, and C5, A6, and C7 are correlated with the C-terminal U1A residues. The residues in the second network are outlined in orange. Tyr13 is strongly correlated with Gln54, Gln54 is correlated to loop 3, especially the N-terminal residues, and the loop 3 residues are strongly to moderately correlated to A1 and the closing C-G base pair of the RNA. Residues in the third network are outlined in yellow. Gly53 is strongly correlated to loop 3, especially the N-terminal residues, loop 3 and loop 1 are strongly correlated, loop 1 is strongly correlated with the 5′ stem and loop nucleotides of RNA, loop 3 is strongly to moderately correlated to A1 and the closing C-G base pair of the RNA, and C5, A6 and C7 are strongly correlated to T89-D90-S91. These results show that residues found to be energetically coupled are fluctuationally correlated, and that every residue involved in each of the three networks is fluctuationally correlated with the subsequent residue(s) in the network. In addition, Fig. 7A shows that the residues involved in these three networks are also fluctuationally correlated with one another, revealing connections between the three individual networks. For example, Tyr13 is correlated with Gln54, (orange box) which is correlated with loop 1 (yellow box). The overall picture emerging from this analysis is that a network of fluctuational correlations can be mapped out connecting all three networks, indicating that these residues are a part of one larger cooperative hypernetwork.
To further explore this idea, Fig. 7B outlines all of the residues implicated in the three cooperative networks and their cross-correlated fluctuations with each other. This analysis indicates that the majority of the residues implicated in the three cooperative networks are fluctuationally correlated with each other with cross-correlation coefficients greater than 0.25 or less than −0.25, supporting the idea that the cooperative networks are all part of a larger hypernetwork of cooperativity. The residues that reveal cross-correlation coefficients between −0.25 – 0.25, indicated by white space, are important for further clarifying the intricacies of the cooperative network. For instance, loop 3 has been implicated in the networks; however all of the residues in loop 3 are not fluctuationally correlated with all of the other residues. Tyr13 is correlated with the Lys50 – Gln54 residues of loop 3, while Thr89 – Ser91 are correlated with the Ser46 – Met51 and Gln 54 residues of loop 3. This DCCM analysis helps to clarify the cooperative network in U1A-RNA through fluctuational correlations. These results, along with those reported by Showalter and Hall (2005) from the analysis of MD using RED, provide encouraging evidence for the use of calculated fluctuational cross-correlations to map out cooperativity in protein-nucleic acid systems.
4.4. Correlated fluctuations and statistical covariance analysis
Independent validation of the link between fluctuational correlations and cooperativity is provided by comparison of our results to an RRM statistical covariance study. Crowder et al. (2001) analyzed the statistical covariance of 330 RRM-containing proteins, including U1A, and were able to subsequently identify a network of covariant amino acid residues present within those RRMs. The results from Crowder et al. compared with the present study are shown in Fig. 8. Positions of residues exhibiting high covariance (> 3.5 σ) are indicated on the DCCM in Fig. 8 by yellow rectangles. The covariant pairs reveal strongly correlated fluctuations from the calculated cross-correlation coefficients. The majority (76%) of the residues shown to covary exhibit fluctuational correlations stronger than 0.50 in the MD simulation. Only one of the covariant pairs exhibited anticorrelated fluctuations, Arg52 – Phe56. This result provides further evidence that a significant amount of information regarding networks of interactions between residues may be extracted from calculation of the cross-correlation coefficients from a MD simulation.
5. Discussion
We have investigated the collective atomic fluctuations of an RRM by calculating the cross-correlation coefficients from a 10 ns MD simulation on the U1A-RNA complex. Analysis was facilitated by displaying the results in the form of a DCCM, which provides a picture of the cross-correlations of atomic fluctuations of U1A-RNA over the entire simulation. Highly correlated fluctuations were found, as expected, between residues and nucleotides involved in contact interactions. More important to this study is that the calculated cross-correlations are indicative of considerable through-space correlation as well, which is not obvious from simply examining the structure per se.
These calculations provide a definition of “dynamical contact” which may carry different information than spatial contacts. Within the protein-protein intramolecular section of the DCCM, well differentiated patterns of positive correlated fluctuations are exhibited that interconnect the entire RRM. This indicates that the RRM of U1A moves in concert as one unit. This observation is consistent with other studies that have shown that U1A has a rigid binding surface (Reyes and Kollman, 1999; Kranz and Hall, 1999; Showalter and Hall, 2005) and that U1A unfolds in a manner consistent with a two state model (Silow and Oliveberg, 1997), however it also shows that concerted fluctuations are exhibited beyond the binding surface and throughout the RRM. The majority of protein-protein anticorrelated cross-correlations seen in the DCCM occur in the N- and C-terminal residues, indicating fluctuations of the flexible N- and C-terminal tails in response to the global motion of the rest of the molecule. The intermolecular fluctuations between protein and RNA that appear from the DCCM analysis indicate correlated motions between the RNA loop nucleotides and the U1A protein due to the stacking interactions with the RRM, and anticorrelated motions between the RNA stem nucleotides and the U1A protein. The anticorrelated fluctuations between U1A and the RNA stem nucleotides indicate a global motion between U1A and RNA, in which the RNA stem and protein move toward each other and then away from each other throughout the MD simulation.
The results of our calculated cross–correlations from MD simulations agree very well with experimentally observed cooperativities in U1A. Moreover, the MD results show cross-correlations predicting that the observed networks are all components of one larger hyper-network. It can be seen from Fig. 7B that this network is quite extensive. However, these residues that have been studied to date by experimental methods only encompass a fraction of the total number of residues in the U1A-RNA, and there are a significant amount of correlated fluctuations in the U1A-RNA system outside of these residues. This indicates the potential for the discovery of additional residues that contribute to cooperativity in this system.
The agreement between our cross-correlation results and cooperativity studies suggests the possibility of using the cross-correlation data to predict additional residues involved in cooperative networks in U1A-RNA. For example, the residues that have cross-correlation values greater than 0.75 with residues that are hydrogen bonded or in VDW contact with RNA nucleotides could be investigated. From such an analysis, several residues have a greater occurrence than the others (Data not shown). These residues include Asp42, Ile43, Leu44, Ala55 and Val57, all of which may contribute to the cooperative network of interactions in U1A-RNA. Of these residues, only Leu44 has been previously studied, the results showing that it is crucial for specificity(Scherly et al., 1990) and that mutation of Leu44 to Val causes a 3.3-fold loss in binding affinity (Rimmele and Belasco, 1998). These findings support the utility of DCCMs for predicting residues important in the network of interactions in U1A-RNA.
It is particularly significant that our calculated cross–correlations are in agreement with results of the statistical covariance analysis of 330 RRMs, which is validation for making additional predictions based on covariance information. For example, several of the covariant residues from this study contact SL2 RNA in the U1A-RNA complex, including Tyr13, Asn15, Asn16, Arg52, Gly53, Gln54 and Phe56. The residues with which each of these residues covary may also be important in the network of interactions in U1A-RNA. The positions of these residues can be seen in Fig. 9. Of these residues, the strongest correlated fluctuations involve Ile14 (with Gln54) and Met82 (with Asn16), moderate correlated fluctuations involve Leu17 (with Asn16), Leu26 (with Asn16), Phe34 (with Asn15), and Met72 (with Asn16), and weaker correlated fluctuations involve Gly38 (with Tyr13). None of these residues have been subjected to mutational analysis to date; however, they show potential to be interesting candidates for the study of cooperative networks in the U1A-RNA system.
Of course, many factors are involved in the binding of any macromolecules, and it is important when investigating binding and recognition to study not only the bound form, but the free constituents as well. Cooperative interactions important for recognition in the unbound macromolecules may change significantly upon binding in order to maintain the complex. As such, dynamical cross-correlation studies on the free constituents and mutants are forthcoming to address issues of affinity and specificity within the U1A-RNA system.
Although a direct connection between fluctuational correlations and cooperativity has not necessarily been proven with this study, the analysis reveals a significant link that should continue to be explored. We plan on further probing this connection between collective atomic fluctuations and cooperativity on small test systems in which cooperativity has been observed. Other potential applications of this method include the study of evolutionarily conserved cooperative networks in RRMs as done for the DNA cytosine methyltransferase family (Estabrook et al., 2005).
6. Summary and Conclusions
We have presented here a dynamical picture of an RRM obtained by calculating the cross-correlations of atomic fluctuations from a MD simulation of the U1A-RNA complex. We have found that cross-correlated fluctuations in the U1A-RNA complex can account for the observed cooperative effects from biophysical experiments and the results of a statistical covariance analysis on 330 examples of biological RRMs. Our calculations suggest that the various observed cooperative networks are correlated with each other, which is consistent with the observed rigidity of the U1A binding surface and the two state folding pathway of U1A. Finally, based on this analysis, we have presented additional residues in U1A that may be important for recognition and binding in this system. The results support the hypothesis that collective fluctuations are indicative of cooperativity. The combination of MD and DCCM analysis is thus a powerful and unique method for understanding and predicting cooperativity at the molecular level in protein-RNA binding, and for elucidating long range, through-space communication in a protein-nucleic acid complex. This approach can be adapted to the study of cooperativity in many systems, including other protein-nucleic acid or protein-protein systems.
Acknowledgments
We would like to thank Dr. Surjit Dixit for many helpful discussions. Funding was provided by the NIH to AMB (GM-56857) and to DLB (GM-076490). BLK was supported by a NIH Postdoctoral Fellowship (F32 GM072345). This work was partially supported by the National Science Foundation through TeraGrid resources provided by Indiana University, the National Center for Supercomputing Applications under LRAC/NRAC grant number MCA94P011 utilizing the Xeon Linux Cluster, and by the Texas Advanced Computing Center utilizing the Cray-Dell PowerEdge Xeon Linux Cluster.
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
References
- Abseher R, Nilges M. Are there non-trivial dynamic cross-correlations in proteins? J Mol Biol. 1998;279:911–920. doi: 10.1006/jmbi.1998.1807. [DOI] [PubMed] [Google Scholar]
- Ackers GK, Smith FR. Effects of Site-Specific Amino Acid Modification on Protein Interactions and Biological Function. Annu Rev Biochem. 1985;54:597–629. doi: 10.1146/annurev.bi.54.070185.003121. [DOI] [PubMed] [Google Scholar]
- Agarwal PK. Cis/trans Isomerization in HIV-1 Capsid Protein Catalyzed by Cyclophilin A: Insights from Computational and Theoretical Studies. Proteins: Struct, Funct, Bioinf. 2004;56:449–463. doi: 10.1002/prot.20135. [DOI] [PubMed] [Google Scholar]
- Allain FH, Howe PW, Neuhaus D, Varani G. Structural basis of the RNA-binding specificity of human U1A protein. EMBO J. 1997;16:5764–5772. doi: 10.1093/emboj/16.18.5764. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arcangeli C, Bizzarri AR, Cannistraro S. Long-term molecular dynamics simulation of copper azurin: structure, dynamics and functionality. Biophys Chem. 1999;78:247–257. doi: 10.1016/s0301-4622(99)00029-0. [DOI] [PubMed] [Google Scholar]
- Arcangeli C, Bizzarri AR, Cannistraro S. Molecular dynamics simulation and essential dynamics study of mutated plastocyanin: structural, dynamical and functional effects of a disulfide bridge insertion at the protein surface. Biophys Chem. 2001;92:183–199. doi: 10.1016/s0301-4622(01)00199-5. [DOI] [PubMed] [Google Scholar]
- Avis JM, Allain FH, Howe PW, Varani G, Nagai K, Neuhaus D. Solution structure of the N-terminal RNP domain of U1A protein: the role of C-terminal residues in structure stability and RNA binding. J Mol Biol. 1996;257:398–411. doi: 10.1006/jmbi.1996.0171. [DOI] [PubMed] [Google Scholar]
- Bahar I, Erman B, Haliloglu T, Jernigan RL. Efficient Characterization of Collective Motions and Interresidue Correlations in Proteins by Low-Resolution Simulations. Biochemisty. 1997;36:13512–13523. doi: 10.1021/bi971611f. [DOI] [PubMed] [Google Scholar]
- Bentley RC, Keene JD. Recognition of U1 and U2 Small Nuclear RNAs Can Be Altered by a 5-Amino-Acid Segment in the U2 Small Nuclear Ribonucleoprotein Particle (snRNP) B” Protein and through Interactions with U2 snRNP-A’ Protein. Mol Cell Biol. 1991;11:1829–1839. doi: 10.1128/mcb.11.4.1829. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Berman HM, Westbrook J, Feng Z, Gilliland G, Bhat TN, Weissig H, Shindyalov IN, Bourne PE. The Protein Data Bank. Nucleic Acids Res. 2000;28:235–242. doi: 10.1093/nar/28.1.235. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Blakaj DM, McConnell KJ, Beveridge DL, Baranger AM. Molecular dynamics and thermodynamics of protein-RNA interactions: Mutation of a conserved aromatic residue modifies stacking interactions and structural adaptation in the U1A-stem loop 2 RNA complex. J Am Chem Soc. 2001;123:2548–2551. doi: 10.1021/ja005538j. [DOI] [PubMed] [Google Scholar]
- Boelens W, Scherly D, Jansen EJR, Kolen K, Mattaj IW, Venrooij WJv. Analysis of in vitro binding of U1-A protein mutants to U1 snRNA. Nucleic Acids Res. 1991;19:4611–4618. doi: 10.1093/nar/19.17.4611. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brooks CL, III, Karplus M, Pettitt BM. Proteins: A Theoretical Perspective of Dynamics, Structure and Thermodynamics. John Wiley and Sons; New York: 1988. [Google Scholar]
- Brünger AT, Brooks CL, Karplus M. Active Site Dynamics of Ribonuclease. Proc Natl Acad Sci USA. 1985;82:8458–8462. doi: 10.1073/pnas.82.24.8458. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Case DA, Darden TA, Cheatham TE, III, Simmerling CL, Wang J, Duke RE, Luo R, Merz KM, Pearlman DA, Crowley M, Walker R, Zhang W, Wang B, Hayik S, Roitberg A, Seabra G, Wong K, Paesani F, Wu X, Brozell S, Tsui V, Gohlke H, Yang L, Tan C, Mongan J, Hornak V, Cui G, Beroza P, Mathews D, Schafmeister C, Ross WS, Kollman P. AMBER 9. University of California; 2006. [Google Scholar]
- Case DA, Darden TA, Cheatham TE, III, Simmerling CL, Wang J, Duke RE, Luo R, Merz KM, Wang B, Pearlman DA, Crowley M, Brozell S, Tsui V, Gohlke H, Mongan J, Hornak V, Cui G, Beroza P, Schafmeister C, Caldwell JW, Ross WS, Kollman P. AMBER 8. University of California; 2004. [Google Scholar]
- Chen Y, Varani G. Protein families and RNA recognition. FEBS J. 2005;272:2088–2097. doi: 10.1111/j.1742-4658.2005.04650.x. [DOI] [PubMed] [Google Scholar]
- Crowder S, Holton J, Alber T. Covariance Analysis of RNA Recognition Motifs Identifies Functionally Linked Amino Acids. J Mol Biol. 2001;310:793–800. doi: 10.1006/jmbi.2001.4740. [DOI] [PubMed] [Google Scholar]
- Di Cera E. Site-Specific Analysis of Mutational Effects in Proteins. Adv Prot Chem. 1998a;51:59–119. doi: 10.1016/s0065-3233(08)60651-8. [DOI] [PubMed] [Google Scholar]
- Di Cera E. Site-Specific Thermodynamics: Understanding Cooperativity in Molecular Recognition. Chem Rev. 1998b;98:1563–1591. doi: 10.1021/cr960135g. [DOI] [PubMed] [Google Scholar]
- Doniach S, Eastman P. Protein dynamics simulations from nanoseconds to microseconds. Curr Opin Struct Biol. 1999;9:157–163. doi: 10.1016/S0959-440X(99)80022-0. [DOI] [PubMed] [Google Scholar]
- Ehresman J, Riemer T, Parker N, Ravishanker G. WesGraph. Wesleyan University; 1995. [Google Scholar]
- Estabrook RA, Luo J, Purdy MM, Sharma V, Weakliem P, Bruice TC, Reich NO. Statistical coevolution analysis and molecular dynamics: Identification of amino acid pairs essential for catalysis. Proc Natl Acad Sci USA. 2005;102:994–999. doi: 10.1073/pnas.0409128102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Freire E. The propagation of binding interactions to remote sites in proteins: Analysis of the binding of the monoclonal antibody D1.3 to lysozyme. Proc Natl Acad Sci USA. 1999;96:10118–10122. doi: 10.1073/pnas.96.18.10118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fuxreiter M, Magyar C, Juhász T, Szeltner Z, Polgár L, Simon I. Flexibility of Prolyl Oligopeptidase: Molecular Dynamics and Molecular Framework Analysis of the Potential Substrate Pathways. Proteins: Struct, Funct, Bioinf. 2005;60:504–512. doi: 10.1002/prot.20508. [DOI] [PubMed] [Google Scholar]
- Gayle AY, Baranger AM. Inhibition of the U1A-RNA complex by an aminoacridine derivative. Bioorg Med Chem Lett. 2002;12:2839–2842. doi: 10.1016/s0960-894x(02)00636-4. [DOI] [PubMed] [Google Scholar]
- Gorfe AA, Caflisch A. Functional Plasticity in the Substrate Binding Site of β-Secretase. Structure. 2005;13:1487–1498. doi: 10.1016/j.str.2005.06.015. [DOI] [PubMed] [Google Scholar]
- Green MR. Biochemical mechanisms of constitutive and regulated pre-mRNA splicing. Annu Rev Cell Biol. 1991;7:559–599. doi: 10.1146/annurev.cb.07.110191.003015. [DOI] [PubMed] [Google Scholar]
- Gunasekaran K, Nussinov R. Modulating Functional Loop Movements: The Role of Highly Conserved Residues in the Correlated Loop Motions. ChemBioChem. 2004;5:224–230. doi: 10.1002/cbic.200300732. [DOI] [PubMed] [Google Scholar]
- Hall KB. Interaction of RNA hairpins with the human U1A N-terminal RNA binding domain. Biochemisty. 1994;33:10076–10088. doi: 10.1021/bi00199a035. [DOI] [PubMed] [Google Scholar]
- Harte WE, Jr, Swaminathan S, Beveridge DL. Molecular Dynamics of HIV-1 Protease. Proteins: Struct, Funct, Genet. 1992;13:175–194. doi: 10.1002/prot.340130302. [DOI] [PubMed] [Google Scholar]
- Harte WE, Jr, Swaminathan S, Mansuri MM, Martin JC, Rosenberg IE, Beveridge DL. Domain Communication in the Dynamical Structure of Human Immunodeficiency Virus 1 Protease. Proc Natl Acad Sci USA. 1990;87:8864–8868. doi: 10.1073/pnas.87.22.8864. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hermann T, Westhof E. Simulations of the dynamics at an RNA-protein interface. Nat Struct Biol. 1999;6:540–544. doi: 10.1038/9310. [DOI] [PubMed] [Google Scholar]
- Hilser VJ, Dowdy D, Oas TG, Freire E. The structural distribution of cooperative interactins in proteins: Analysis of the native state ensemble. Proc Natl Acad Sci USA. 1998;95:9903–9908. doi: 10.1073/pnas.95.17.9903. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hornak V, Abel R, Okur A, Strockbine B, Roitberg A, Simmerling C. Comparison of multiple Amber force fields and development of improved protein backbone parameters. 2006. Submitted for publication. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Horovitz A, Fersht AR. Strategy for Analysing the Co-operativity of Intramolecular Interactions in Peptides and Proteins. J Mol Biol. 1990;214:613–617. doi: 10.1016/0022-2836(90)90275-Q. [DOI] [PubMed] [Google Scholar]
- Humphrey W, Dalke A, Schulten K. VMD - (V)isual (M)olecular (D)ynamics. J Molec Graphics. 1996;14:33–38. doi: 10.1016/0263-7855(96)00018-5. [DOI] [PubMed] [Google Scholar]
- Hünenberger PH, Mark AE, van Gunsteren WF. Fluctuation and cross-correlation analysis of protein motions observed in nanosecond molecular dynamics simulations. J Mol Biol. 1995;252:492–503. doi: 10.1006/jmbi.1995.0514. [DOI] [PubMed] [Google Scholar]
- Ichiye T, Karplus M. Collective Motions in Proteins: A Covariance Analysis of Atomic Fluctuations in Molecular Dynamics and Normal Mode Simulations. Proteins: Struct, Funct, Genet. 1991;11:205–217. doi: 10.1002/prot.340110305. [DOI] [PubMed] [Google Scholar]
- Jessen TH, Oubridge C, Teo CH, Pritchard C, Nagai K. Identification of molecular contacts between the U1 A small nuclear ribonucleoprotein and U1 RNA. EMBO J. 1991;10:3447–3456. doi: 10.1002/j.1460-2075.1991.tb04909.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jorgensen WL, Chandrasekhar J, Madura JD, Impey RW, Klein ML. Comparison of Simple Potential Functions for Simulating Liquid Water. J Chem Phys. 1983;79:926–936. [Google Scholar]
- Karplus M, Ichiye T. Comment on a “fluctuation and cross correlation analysis of protein motions observed in nanosecond molecular dynamics simulations”. J Mol Biol. 1996;263:120–122. doi: 10.1006/jmbi.1996.0562. [DOI] [PubMed] [Google Scholar]
- Karplus M, Kuriyan J. Molecular dynamics and protein function. Proc Natl Acad Sci USA. 2005;102:6679–6685. doi: 10.1073/pnas.0408930102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Karplus M, McCammon JA. Molecular dynamics simulations of biomolecules. Nat Struct Biol. 2002;9:646–652. doi: 10.1038/nsb0902-646. [DOI] [PubMed] [Google Scholar]
- Karplus M, Petsko GA. Molecular dynamics simulations in biology. Nature. 1990;347:631–639. doi: 10.1038/347631a0. [DOI] [PubMed] [Google Scholar]
- Katsamba PS, Bayramyan M, Haworth IS, Myszka DG, Laird-Offringa IA. Complex role of the β2-β3 Loop in the Interaction of U1A with U1 Hairpin II RNA. J Biol Chem. 2002;36:33267–33274. doi: 10.1074/jbc.M200304200. [DOI] [PubMed] [Google Scholar]
- Kollman PA, Dixon R, Cornell W, Fox T, Chipot C, Pohorille A. The development/application of a ‘minimalist’ organic/biochemicl molecular mechanic force field using a combination of ab initio calculations and experimental data. In: Wilkinson A, Weiner P, van Gunsteren WF, editors. Computer Simulation of Biomolecular Systems. Kluwer; London: 1997. pp. 83–96. [Google Scholar]
- Kranz JK, Hall KB. RNA binding mediates the local cooperativity between the beta-sheet and the C-terminal tail of the human U1A RBD1 protein. J Mol Biol. 1998;275:465–481. doi: 10.1006/jmbi.1997.1441. [DOI] [PubMed] [Google Scholar]
- Kranz JK, Hall KB. RNA recognition by the human U1A protein is mediated by a network of local cooperative interactions that create the optimal binding surface. J Mol Biol. 1999;285:215–231. doi: 10.1006/jmbi.1998.2296. [DOI] [PubMed] [Google Scholar]
- Laird-Offringa IA, Belasco JG. Analysis of RNA-binding proteins by in vitro genetic selection: identification of an amino acid residue important for locking U1A onto its RNA target. Proc Natl Acad Sci USA. 1995;92:11859–11863. doi: 10.1073/pnas.92.25.11859. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lange OF, Grubmüller H, de Groot BL. Molecular Dynamics Simulations of Protein G Challenge NMR-Derived Correlated Backbone Motions. Angew Chem. 2005;44:3394–3399. doi: 10.1002/anie.200462957. [DOI] [PubMed] [Google Scholar]
- Law MJ, Chambers EJ, Katsamba PS, Haworth IS, Laird-Offringa IA. Kinetic analysis of the role of the tyrosine 13, phenylalanine 56 and glutamine 54 network in the U1A/U1 hairpin II interaction. Nucleic Acids Res. 2005;33:2917–2928. doi: 10.1093/nar/gki602. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lockless SW, Ranganathan R. Evolutionarily Conserved Pathways of Energetic Connectivity in Protein Families. Science. 1999;286:295–299. doi: 10.1126/science.286.5438.295. [DOI] [PubMed] [Google Scholar]
- Luchansky SJ, Nolan SJ, Baranger AM. Contribution of RNA Conformation to the Stability of a High-Affinity RNA-Protein Complex. J Am Chem Soc. 2000;122:7130–7131. [Google Scholar]
- Luo J, Bruice TC. Ten-nanosecond molecular dynamics simulation of the motions of the horse liver alcohol dehydrogenase·PhCH2O- complex. Proc Natl Acad Sci USA. 2002;99:16597–16600. doi: 10.1073/pnas.262667599. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Luque I, Leavitt SA, Freire E. The Linkage Between Protein Folding and Functional Cooperativity. Annu Rev Biophys Biomol Struct. 2002;31:235–256. doi: 10.1146/annurev.biophys.31.082901.134215. [DOI] [PubMed] [Google Scholar]
- Ma W, Tang C, Lai L. Specificity of Trypsin and Chymotrypsin: Loop-Motion-Controlled Dynamic Correlation as a Determinant. Biophys J. 2005;89:1183–1193. doi: 10.1529/biophysj.104.057158. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maris C, Dominguez C, Allain FHT. The RNA recognition motif, a plastic RNA-binding platform to regulate post-transcriptional gene expression. FEBS J. 2005;272:2118–2131. doi: 10.1111/j.1742-4658.2005.04653.x. [DOI] [PubMed] [Google Scholar]
- McCammon AJ, Harvey SC. Dynamics of Proteins and Nucleic Acids. Cambridge University Press; Cambridge: 1986. [Google Scholar]
- McCammon JA. Protein Dynamics. Rept Prog Phys. 1984;47:1–46. [Google Scholar]
- Mittermaier A, Varani L, Muhandiram DR, Kay LE, Varani G. Changes in side-chain and backbone dynamics identify determinants of specificity in RNA recognition by human U1A protein. J Mol Biol. 1999;294:967–979. doi: 10.1006/jmbi.1999.3311. [DOI] [PubMed] [Google Scholar]
- Murphy KP, Freire E. Thermodynamics of Structural Stability and Cooperative Folding Behavior in Proteins. Adv Prot Chem. 1992;43:313–361. doi: 10.1016/s0065-3233(08)60556-2. [DOI] [PubMed] [Google Scholar]
- Nolan SJ, Shiels JC, Tuite JB, Cecere KL, Baranger AM. Recognition of an Essential Adenine at a Protein-RNA Interface: Comparison of Hydrogen Bonds and a Stacking Interaction. J Am Chem Soc. 1999;121:8951–8952. [Google Scholar]
- Okur A, Strockbine B, Hornak V, Simmerling C. Using PC Clusters to Evaluate the Transferability of Molecular Mechanics Force Fields for Proteins. J Comput Chem. 2002;24:21–31. doi: 10.1002/jcc.10184. [DOI] [PubMed] [Google Scholar]
- Oubridge C, Ito N, Evans PR, Teo CH, Nagai K. Crystal structure at 1.92 A resolution of the RNA-binding domain of the U1A spliceosomal protein complexed with an RNA hairpin. Nature. 1994;372:432–438. doi: 10.1038/372432a0. [DOI] [PubMed] [Google Scholar]
- Parchment OG, Essex JW. Molecular Dynamics of Mouse and Syrian Hamster PrP: Implications for Activity. Proteins: Struct, Funct, Genet. 2000;38:327–340. [PubMed] [Google Scholar]
- Pearlman DA, Case DA, Caldwell JW, Ross WS, Cheatham TE, DeBolt S, Ferguson D, Seibel G, Kollman P. AMBER, a package of computer programs for applying molecular mechanics, normal mode analysis, molecular dynamics and free energy calculations to simulate the structural and energetic properties of molecules. Comput Phys Commun. 1995;91:1–41. [Google Scholar]
- Persistence of Vision Raytracer. Williamstown, Victoria, Australia. 2004 http://www.povray.org/
- Pitici F, Beveridge DL, Baranger AM. Molecular dynamics simulation studies of induced fit and conformational capture in U1A-RNA binding: Do molecular substates code for specificity? Biopolymers. 2002;65:424–435. doi: 10.1002/bip.10251. [DOI] [PubMed] [Google Scholar]
- Prompers JJ, Brüschweiler R. Dynamic and Structural Analysis of Isotropically Distributed Molecular Ensembles. Proteins: Struct, Funct, Genet. 2002;46:177–189. doi: 10.1002/prot.10025. [DOI] [PubMed] [Google Scholar]
- Radkiewicz JL, Brooks CL., III Protein Dynamics in Enzymatic Catalysis: Exploration of Dihydrofolate Reductase. J Am Chem Soc. 2000;122:225–231. [Google Scholar]
- Ravishanker G, Wang W, Beveridge DL. Molecular Dynamics Tool Chest 2.0. Wesleyan University; 1998. [Google Scholar]
- Reyes CM, Kollman PA. Molecular dynamics studies of U1A-RNA complexes. RNA. 1999;5:235–244. doi: 10.1017/s1355838299981657. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rimmele ME, Belasco JG. Target discrimination by RNA-binding proteins: role of the ancillary protein U2A’ and a critical leucine residue in differentiating the RNA-binding specificity of spliceosomal proteins U1A and U2B”. RNA. 1998;4:1386–1396. doi: 10.1017/s1355838298981171. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rizzuti B, Sportelli L, Guzzi R. Evidence of reduced flexibility in disulfide bridge-depleted azurin: a molecular dynamics simulation study. Biophys Chem. 2001;94:107–120. doi: 10.1016/s0301-4622(01)00228-9. [DOI] [PubMed] [Google Scholar]
- Rod TH, Radkiewicz JL, Brooks CL., III Correlated motion and the effect of distal mutations in dihydrofolate reductase. Proc Natl Acad Sci USA. 2003;100:6980–6985. doi: 10.1073/pnas.1230801100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Scherly D, Boelens W, Dathan NA, van Venrooij WJ, Mattaj IW. Major determinants of the specificity of interaction between small nuclear ribonucleoproteins U1A and U2B” and their cognate RNAs. Nature. 1990;345:502–506. doi: 10.1038/345502a0. [DOI] [PubMed] [Google Scholar]
- Schieborr U, Rüterjans H. Bias-Free Separation of Internal and Overall Motion of Biomolecules. Proteins: Struct, Funct, Genet. 2001;45:207–218. doi: 10.1002/prot.1141. [DOI] [PubMed] [Google Scholar]
- Schiøtt B. Possible Involvement of Collective Domain Movement in the Catalytic Reaction of Soluble Epoxide Hydrolase. Int J Quantum Chem. 2004;99:61–69. [Google Scholar]
- Shiels JC, Tuite JB, Nolan SJ, Baranger AM. Investigation of a conserved stacking interaction in target site recognition by the U1A protein. Nucleic Acids Res. 2002;30:550–558. doi: 10.1093/nar/30.2.550. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Showalter SA, Hall KB. A functional role for correlated motion in the N-terminal RNA-binding domain of human U1A protein. J Mol Biol. 2002;322:533–542. doi: 10.1016/s0022-2836(02)00804-5. [DOI] [PubMed] [Google Scholar]
- Showalter SA, Hall KB. Altering the RNA-binding mode of the U1A RBD1 protein. J Mol Biol. 2004;335:465–480. doi: 10.1016/j.jmb.2003.10.055. [DOI] [PubMed] [Google Scholar]
- Showalter SA, Hall KB. Correlated Motions in the U1 snRNA Stem/Loop 2:U1A RBD1 Complex. Biophys J. 2005;89:2046–2058. doi: 10.1529/biophysj.104.058032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Silow M, Oliveberg M. High-Energy Channeling in Protein Folding. Biochemisty. 1997;36:7633–7637. doi: 10.1021/bi970210x. [DOI] [PubMed] [Google Scholar]
- Stark H, Dube P, Lührmann R, Kastner B. Arrangement of RNA and proteins in the spliceosomal U1 small nuclear ribonucleoprotein particle. Nature. 2001;409:539–542. doi: 10.1038/35054102. [DOI] [PubMed] [Google Scholar]
- Suenaga A, Yatsu C, Komeiji Y, Uebayasi M, Meguro T, Yamato I. Molecular dynamics simulation of trp-repressor/operator complex: analysis of hydrogen bond patterns of protein-DNA interaction. J Mol Struct. 2000;526:209–218. [Google Scholar]
- Sulpizi M, Rothlisberger U, Carloni P. Molecular Dynamics Studies of Caspase-3. Biophys J. 2003;84:2207–2215. doi: 10.1016/S0006-3495(03)75026-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Swaminathan S, Harte WE, Beveridge DL. Investigation of Domain Structure in Proteins via Molecular Dynamics Simulation: Application to HIV-1 Protease Dimer. J Am Chem Soc. 1991;113:2717–2721. [Google Scholar]
- Tang Y, Nilsson L. Molecular dynamics simulations of the complex between human U1A protein and hairpin II of U1 small nuclear RNA and of free RNA in solution. Biophys J. 1999;77:1284–1305. doi: 10.1016/S0006-3495(99)76979-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Trylska J, Tozzini V, McCammon AJ. Exploring Global Motions and Correlations in the Ribosome. Biophys J. 2005;89:1455–1463. doi: 10.1529/biophysj.104.058495. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tsai DE, Harper DS, Keene JD. U1-snRNP-A selects a ten nucleotide consensus sequence from a degenerate RNA pool presented in various structural contexts. Nucleic Acids Res. 1991;19:4931–4936. doi: 10.1093/nar/19.18.4931. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tuite JB, Shiels JC, Baranger AM. Substitution of an essential adenine in the U1A-RNA complex with a non-polar isostere. Nucleic Acids Res. 2002;30:5269–5275. doi: 10.1093/nar/gkf636. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Varani G, Nagai K. RNA recognition by RNP proteins during RNA processing. Annu Rev Biophys Biomol Struct. 1998;27:407–445. doi: 10.1146/annurev.biophys.27.1.407. [DOI] [PubMed] [Google Scholar]
- Wang J, Cieplak P, Kollman PA. How Well Does a Restrained Electrostatic Potential (RESP) Model Perform in Calculating Conformational Energies of Organic and Biological Molecules? J Comput Chem. 2000;21:1049–1074. [Google Scholar]
- Wong KF, Watney JB, Hammes-Schiffer S. Analysis of Electrostatics and Correlated Motions for Hydride Transfer in Dihydrofolate Reductase. J Phys Chem B. 2004;108:12231–12241. [Google Scholar]
- Wyman J. Heme Proteins. Adv Prot Chem. 1948;4:407–531. doi: 10.1016/s0065-3233(08)60011-x. [DOI] [PubMed] [Google Scholar]
- Wyman J, Gill SJ. Binding and Linkage. University Science Books; Mill Valley, CA: 1990. [Google Scholar]
- Young MA, Gonfloni S, Superti-Furga G, Roux B, Kuriyan J. Dynamic Coupling between the SH2 and SH3 Domains of c-Src and Hck Underlies Their Inactivation by C-Terminal Tyrosine Phosphorylation. Cell. 2001;105:115–126. doi: 10.1016/s0092-8674(01)00301-4. [DOI] [PubMed] [Google Scholar]
- Zeng Q, Hall KB. Contribution of the C-terminal tail of U1A RBD1 to RNA recognition and protein stability. RNA. 1997;3:303–314. [PMC free article] [PubMed] [Google Scholar]
- Zhao Y, Baranger AM. Design of an adenosine analogue that selectively improves the affinity of a mutant U1A protein for RNA. J Am Chem Soc. 2003;125:2480–2488. doi: 10.1021/ja021267w. [DOI] [PubMed] [Google Scholar]
- Zhao Y, Kormos BL, Beveridge DL, Baranger AM. Molecular Dynamics Simulation Studies of a Protein-RNA Complex with a Selectively Modified Binding Interface. Biopolymers. 2006;81:256–269. doi: 10.1002/bip.20408. [DOI] [PubMed] [Google Scholar]
- Zhou Y, Cook M, Karplus M. Protein motions at zero-total angular momentum: the importance of long-range correlations. Biophys J. 2000;79:2902–2908. doi: 10.1016/S0006-3495(00)76527-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zoete V, Michielin O, Karplus M. Relation between Sequence and Structure of HIV-1 Protease Inhibitor Complexes: A Model System for the Analysis of Protein Flexibility. J Mol Biol. 2002;315:21–52. doi: 10.1006/jmbi.2001.5173. [DOI] [PubMed] [Google Scholar]