SMB | SMT | SMRA | Main Page

 

INTERMOLECULAR FORCES, RECOGNITION, AND DYNAMICS

 

Donald C. Rau, PhD, Head, Section on Molecular Recognition and Assembly

Nina Sidorova, PhD, Staff Fellow

Brian Todd, PhD, Postdoctoral Fellow

Shakir Muradymov, Summer Intern

 

 

 

 

 

 

 

Our laboratory focuses on elucidating the coupling of the forces, structure, and dynamics of biologically important macromolecules. The next challenge in structural biology is to understand the physics of interactions between molecules in aqueous solution. The ability to take advantage of the increasing number of protein and nucleic acid structures determined by X-ray crystallography and solution NMR will depend critically on knowledge of structural biology so that we can understand the strength and specificity of interactions among biologically important macromolecules that control cellular function and then rationally design agents that can effectively compete with those specific interactions associated with disease. Our earlier results have shown that experimentally measured forces differ markedly from those predicted by current, conventionally accepted theories of intermolecular interactions. We have interpreted the observed forces as indicating the dominant contribution of water-structuring energetics. Our research program uses osmotic stress and X-ray scattering to measure directly forces between biological macromolecules in macroscopic condensed arrays. To investigate the role of water in the interaction of individual molecules, we measure and correlate changes in binding energies as well as in the hydration accompanying specific recognition reactions of biologically important macromolecules in dilute solution, particularly of sequence-specific DNA-protein complexes.

1. Direct Force Measurements

The ability to measure directly forces between biopolymers in macroscopic condensed arrays has greatly changed our understanding of how molecules interact at close spacings, i.e., at 10 to 15Å separation. The universality of the force characteristics observed for a wide variety of macromolecules, including DNA, proteins, lipid bilayers, and carbohydrates, has led us to conclude that the energy associated with structuring water between close surfaces dominates intermolecular forces. We are currently focusing on understanding the connection between hydration force magnitudes and the chemical natures of the interacting surfaces.

Exclusion of solutes from macromolecular surfaces

Rau

Using direct force measurement, we have further characterized the exclusion of alcohols from DNA. Forces between DNA helices in macroscopic condensed arrays can be measured by monitoring the distance between helices, using X-ray diffraction as a function of the osmotic pressure of a polymer solution in equilibrium with the DNA phase. The sensitivity of DNA forces to the concentration of alcohols or other solutes permits extraction of the spatial distribution of alcohols or other solutes around DNA. We have previously seen that repulsive hydration forces underlie the preferential hydration of DNA in the presence of alcohols. We have observed little difference in the spatial dependence associated with the exclusion of methyl pentanediol (MPD) from spermidine (Spd3+)-DNA, Co3+-DNA, or NaBr-DNA arrays. The hydration of phosphate groups on the DNA backbone likely dominates forces. Any additional contribution from electrostatics and a lowered solution dielectric constant is minimal.

To analyze the chemical features of alcohols that determine exclusion, we are examining the interaction of 17 alcohols with Spd3+-DNA arrays. To a surprisingly good first-order approximation, the amplitudes of the observed repulsive hydration forces simply scale with the number of alkyl carbons in the alcohol. In contrast, carbons with bound hydroxyl groups are nearly invisible to DNA. Methanol and ethylene glycol are very slightly excluded from Spd3+-DNA arrays; glycerol is neither included nor excluded; threitol and sorbitol are slightly included, as is the six-carbon polyol sorbitol, more so than threitol with four hydroxylcarbons. Our current hypothesis posits that these polyols can replace strings of water bound to DNA with a subsequent gain in entropy.

Hultgren A, Rau DC. Exclusion of alcohols from spermidine-DNA assemblies: probing the physical basis of preferential hydration. Biochemistry 2004;43:8272-8280.

Spermidine-DNA assemblies

Rau; in collaboration with Yang

We have completed our investigation into the unusual spermidine-mediated transition between two attractive states of DNA, a hexagonal phase seen at lower spermidine (Spd) concentrations and a cholesteric phase with a larger interhelical spacing that appears at higher Spd concentrations. The transition has been widely interpreted as overcharging of DNA, i.e., the binding of more Spd than is necessary to neutralize the DNA charge. However, careful examination of the dependence of the critical Spd concentration at the transition point on NaCl concentration shows that overcharging does not occur. The data strongly suggest that, at higher Spd concentrations, the bulk solution contains a mixture of Spd3+ and SpdCl2+ ions. The data are best fit with a Cl binding constant of about 100 mM. The structural transition is most likely attributable to a replacement of trivalent Spd with a more weakly attractive or perhaps even repulsive divalent Spd-Cl complex.

Single-molecule force measurements

Todd, Rau; in collaboration with Parsegian

We have finished constructing a magnetic tweezers apparatus for single-molecule force measurements. Our initial experiments will focus on measuring the interhelical attractive energies for DNA condensed with various divalent, trivalent, and higher-valence cation ions, including naturally occurring protamines used for tightly packaging sperm DNA in several organisms. The goal is to connect the dependence of attractive energies on equilibrium interhelical spacing to hydration forces.

2. Hydration Changes Linked to Sequence-Specific DNA-Protein Recognition Reactions

Our ultimate goal is to apply the lessons from direct force measurements to the recognition reactions that control cellular processes. We have started by measuring differences in water sequestered by complexes of four sequence-specific DNA binding proteins with varying DNA sequences, with particular emphasis on correlating the incorporated binding energy and water as well as on the energy necessary to remove hydrating water from complexes.

The energetics of removing water from noncognate EcoRI-DNA complexes

Sidorova, Rau

We have continued investigating the role of water in the sequence-specific recognition reaction of the restriction nuclease EcoRI. The exceptionally stringent specificity of restriction endonucleases in general and of the EcoRI in particular is a paradigm for DNA-protein recognition reactions. We measure the number of waters coupled to the DNA-protein binding reaction from the dependence of the binding free energy on water chemical potential or, equivalently, on osmotic pressure. We previously found that the nonspecific complex of the EcoRI sequesters about 110 more waters than the complex with the specific recognition sequence and that the water is likely located at the protein-DNA interface. In contrast to this extremely strong sensitivity to the osmotic stress (or water activity), the relative specific-nonspecific binding constant has a particularly weak salt dependence and no pH dependence. We have further shown that the dissociation rate of the specific complex is highly sensitive to osmotic pressure requiring the net binding of 120 to 150 waters for the dissociation reaction.

During the past year, we completed our project on the dependence of the number of waters sequestered on DNA sequence. Sequences that differ by even a single base pair (“star” sequences) from the EcoRI-specific recognition sequence (GAATTC) bind to protein only slightly better than completely nonspecific sequences. We demonstrated that this abrupt decrease in binding energy with even a single base-pair change is accompanied by an abrupt increase in the water sequestered by the EcoRI-DNA complex. We developed a new quantitative technique allowing us to measure reliably the dissociation rate of the EcoRI from noncognate DNA sequences. At low stresses, we observed that the osmotic sensitivities of both the competitive equilibrium constants and dissociation rates of complexes with two “star” sequences that differ by a single base pair out of six from the specific recognition sequence and with nonspecific DNA are practically indistinguishable. At low osmotic stresses, the complexes with noncognate sequences sequestered about the same 110 waters as nonspecific complexes. At least some of the water, however, can be removed from the noncognate complexes by applying sufficiently high osmotic pressures. Given that equilibration times at high stresses are too long for convenient measurement of binding constants, we used dissociation rates to determine loss of sequestered water. At higher pressures, the behavior of “star” sequence complexes differed strikingly from that of the nonspecific complex. We observed a clear correlation between the osmotic dependence of the dissociation rate and DNA sequence. The kinetic measurements showed that the TAATTC “star” sequence complex (strongest “star” site for the EcoRI) loses about 90 of its initially sequestered waters at an osmotic pressure of 150 atmospheres. The osmotic work required to remove these waters was about 3 Kcal/mole. Over the same range of pressures, the CAATTC complex, the weaker “star” site, lost only about 20 to 30 of its waters; we estimated that about 6 Kcal/mole would be necessary to remove substantially all the water from this complex. There was no apparent loss of water from the nonspecific complex. The amount of water we were able to remove from a noncognate complex correlated with the ability of the enzyme to cleave the sequence. In principle, any sequestered water can be removed by applying sufficiently high osmotic stress, but the work necessary to dehydrate complexes naturally depends on resulting unfavorable DNA-protein contacts.

Sidorova NY, Rau DC. Differences between EcoRI nonspecific and “star” sequence complexes revealed by osmotic stress. Biophys J 2004;87:2564-2576.

Differences in sequestered water between specific and nonspecific complexes of BamHI: comparison with X-ray structures

Muradymov, Sidorova, Rau

We previously observed that the osmotic sensitivity of the EcoRI specific-nonspecific equilibrium binding constant was only very slightly dependent on the nature of the solute used to set water activity. That observation led us to conclude that the extra water retained by the nonspecific complex is sequestered in a space sterically inaccessible to solutes, most probably at the DNA-protein interface. As no structure of the nonspecific complex of EcoRI is available, we could not directly confirm our conclusion. To validate the connection between osmotic stress measurements and structure, we are now investigating the difference in water release and DNA binding strength for a second restriction endonuclease, Bam HI. The significant advantage of this system is that X-ray structures are available for both specific and nonspecific BamHI-DNA complexes. A comparison of the two structures showed that the nonspecific complex loses the direct contacts of the protein with DNA bases in the major groove that is characteristic of the specific sequence complex. While extensive contacts still exist between the sugar-phosphate backbone and the protein, a large cavity separates the major groove of the DNA and the recognition protein surface. We estimated the volume of the cavity as equivalent to about 150 waters. We performed a series of kinetic and equilibrium competition experiments, initially with a commercially available protein and subsequently with a highly purified sample of the enzyme. For both protein samples, we observed very strong osmotic dependence of the specific dissociation rate constant, with half-life time changing from two minutes with no osmolyte present to about 40 minutes in the presence of 0.8 osmolal betaine or methyl glucoside. Initial equilibrium competition measurements indicated that the nonspecific BamHI-DNA complex sequesters about 130 waters more than the specific complex of the enzyme, correlating well with the structural data.

Sidorova NY, Rau DC. The role of water in EcoRI-DNA binding. In: Pingoud A, ed. Restriction Endonucleases, Nucleic Acids and Molecular Biology, Volume 14. Berlin: Springer, 2004;319-337.

3. Hydration Changes Associated with Specific Lambda Cro-DNA Binding

We are continuing to investigate the correlation between water and binding strength of DNA complexes with lambda Cro repressor. This protein is responsible for the induction of the lytic phase of lambda phage. The recognition stringency for the repressor is more typical of sequence-specific DNA-protein binding systems. Single base-pair changes from the optimal recognition sequence reduce binding energies by only a few kT. We have constructed a series of DNA fragments with consensus and altered Cro recognition sequences that span a range of about 1,000 in binding constant. Results for the osmotic pressure dependence of dissociation rates and of the competitive equilibrium binding constants for the various complexes indicated that the number of sequestered waters increases as the binding energy decreases. For about every 10-fold decrease in the binding constant, the complex incorporates an extra 10 to 15 water molecules. The insensitivity of the number of extra water molecules sequestered by these complexes to solute identity indicated that the waters are also likely at the DNA-protein interface, mediating suboptimal contacts between the two surfaces.

Sensitivity of Thermus aquaticus MutS-DNA–specific binding to solution conditions

Sidorova

The mismatch repair system, which is highly conserved in both prokaryotes and eukaryotes, plays a crucial role in maintaining genetic stability. Among a large group of proteins required for mismatch repair, only MutS recognizes and binds to heteroduplex DNA containing mispaired or unpaired bases. MutS binding then triggers a cascade of events ultimately leading to repair of the mismatched sequence. All members of MutS family possess ATPase activity, and both mismatch recognition and the ATPase activity of MutS are critical for the overall performance of the mismatch repair system. However, it is still unclear how MutS recognizes a broad range of mismatches (one to four consecutive unpaired bases or mispaired bases such as GT, AG, and so forth). Broad substrate specificity is typical for the proteins of the MutS family. The binding preferences and specificities of MutS proteins differ markedly from the sequence-specific DNA-binding proteins such as restriction endonucleases or lambda Cro repressor that we have previously studied.

The structure of Taq MutS in complex with DNA was recently solved. Both biochemical studies and the crystal structure indicated that the MutS dimer is responsible for specific mismatch recognition. Dimerization of the protein, which buries about 2,400 Å2 of molecular surface, extensive interactions between kinked heteroduplex DNA, and four MutS domains plus the presence of an extra channel, whose function in the MutS-DNA complex structure is still unknown, suggest that MutS-DNA binding could be osmotically sensitive. We are now investigating the effect of solution thermodynamic parameters (salt activity, water activity, and pH) for mismatch recognition by MutS.

We employed gel mobility shift assays to study DNA binding of Taq MutS protein to two DNA heteroduplexes containing one unpaired base each. Binding curves of MutS to a 60 bp DNA heteroduplex containing unpaired C have a distinct sigmoidal shape, typical of positively cooperative binding. Scatchard plots confirmed that MutS binding to this DNA substrate is highly cooperative. We suggest that observed positive cooperativity results from monomer-dimer equilibrium coupled with binding of the dimeric form of Taq MutS to DNA. A theoretical model that accounts for both processes allows us to quantitate the apparent binding constant and estimate its sensitivity to solution conditions. In parallel with gel mobility shift experiments, we are also employing fluorescence anisotropy to measure MutS-DNA binding. Dissociation kinetic experiments performed with no osmolyte and in the presence of one osmolal betaine glycine or triethylene glycol showed that the stability of the MutS-DNA specific complex increases significantly in the presence of solutes with the half-life time increasing about 9-fold between zero and one osmolal betaine or triethylene glycol.

COLLABORATORS

Per Lyngs Hansen, PhD, University of Southern Denmark, Odense, Denmark

Peggy Hsieh, PhD, Genetics and Biochemistry Branch, NIDDK, Bethesda, MD

Sergey Leikin, PhD, Section on Molecular Forces and Assembly, OD, NICHD, Bethesda, MD

Shimon Mizrahi, PhD, Technion-Israel Institute of Technology, Haifa, Israel

Galina Obmolova, PhD, Genetics and Biochemistry Branch, NIDDK, Bethesda, MD

V. Adrian Parsegian, PhD, Laboratory of Physical and Structural Biology, NICHD, Bethesda, MD

Rudi Podgornik, PhD, Laboratory of Physical and Structural Biology, NICHD, Bethesda, MD

Jie Yang, PhD, University of Vermont, Burlington, VT


For further information, contact donrau@helix.nih.gov