Publications

2013
J. K. Fisher, A. Bourniquel, G. Witz, B. Weiner, M. Prentiss, and N. Kleckner. 2013. “Four-Dimensional Imaging of E. coli Nucleoid Organization and Dynamics in Living Cells.” CELL, 153, Pp. 882-895. Publisher's VersionAbstract
Visualization of living E. coli nucleoids, defined by HupA-mCherry, reveals a discrete, dynamic helical ellipsoid. Three basic features emerge. (1) Nucleoid density coalesces into longitudinal bundles, giving a stiff, low-DNA-density ellipsoid. (2) This ellipsoid is radially confined within the cell cylinder. Radial confinement gives helical shape and directs global nucleoid dynamics, including sister segregation. (3) Longitudinal density waves flux back and forth along the nucleoid, with 5%-10% of density shifting within 5 s, enhancing internal nucleoid mobility. Furthermore, sisters separate end-to-end in sequential discontinuous pulses, each elongating the nucleoid by 5%-15%. Pulses occur at 20 min intervals, at defined cell-cycle times. This progression includes sequential installation and release of programmed tethers, implying cyclic accumulation and relief of intranucleoid mechanical stress. These effects could comprise a chromosome-based cell-cycle engine. Overall, the presented results suggest a general conceptual framework for bacterial nucleoid nnorphogenesis and dynamics.
J. Kates-Harbeck, A. Tilloy, and M. Prentiss. 2013. “Simplified biased random walk model for RecA-protein-mediated homology recognition offers rapid and accurate self-assembly of long linear arrays of binding sites.” PHYSICAL REVIEW E, 88. Publisher's VersionAbstract
Inspired by RecA-protein-based homology recognition, we consider the pairing of two long linear arrays of binding sites. We propose a fully reversible, physically realizable biased random walk model for rapid and accurate self-assembly due to the spontaneous pairing of matching binding sites, where the statistics of the searched sample are included. In the model, there are two bound conformations, and the free energy for each conformation is a weakly nonlinear function of the number of contiguous matched bound sites.
J. Vlassakis, E. Feinstein, D. Yang, A. Tilloy, D. Weiller, J. Kates-Harbeck, V. Coljee, and M. Prentiss. 2013. “Tension on dsDNA bound to ssDNA-RecA filaments may play an important role in driving efficient and accurate homology recognition and strand exchange.” PHYSICAL REVIEW E, 87. Publisher's VersionAbstract
It is well known that during homology recognition and strand exchange the double-stranded DNA (dsDNA) in DNA/RecA filaments is highly extended, but the functional role of the extension has been unclear. We present an analytical model that calculates the distribution of tension in the extended dsDNA during strand exchange. The model suggests that the binding of additional dsDNA base pairs to the DNA/RecA filament alters the tension in dsDNA that was already bound to the filament, resulting in a nonlinear increase in the mechanical energy as a function of the number of bound base pairs. This collective mechanical response may promote homology stringency and underlie unexplained experimental results. DOI: 10.1103/PhysRevE.87.032702
2012
S. Tricard, E. Feinstein, R. F. Shepherd, M. Reches, P. W. Snyder, D. C. Bandarage, M. Prentiss, and G. M. Whitesides. 2012. “Analog modeling of Worm-Like Chain molecules using macroscopic beads-on-a-string.” PHYSICAL CHEMISTRY CHEMICAL PHYSICS, 14, Pp. 9041-9046. Publisher's VersionAbstract
This paper describes an empirical model of polymer dynamics, based on the agitation of millimeter-sized polymeric beads. Although the interactions between the particles in the macroscopic model and those between the monomers of molecular-scale polymers are fundamentally different, both systems follow the Worm-Like Chain theory.
A. Peacock-Villada, D. Yang, C. Danilowicz, E. Feinstein, N. Pollock, S. McShan, V. Coljee, and M. Prentiss. 2012. “Complementary strand relocation may play vital roles in RecA-based homology recognition.” NUCLEIC ACIDS RESEARCH, 40, Pp. 10441-10451. Publisher's VersionAbstract
RecA-family proteins mediate homologous recombination and recombinational DNA repair through homology search and strand exchange. Initially, the protein forms a filament with the incoming single-stranded DNA (ssDNA) bound in site I. The RecA-ssDNA filament then binds double-stranded DNA (dsDNA) in site II. Non-homologous dsDNA rapidly unbinds, whereas homologous dsDNA undergoes strand exchange yielding heteroduplex dsDNA in site I and the leftover outgoing strand in site II. We show that applying force to the ends of the complementary strand significantly retards strand exchange, whereas applying the same force to the outgoing strand does not. We also show that crystallographically determined binding site locations require an intermediate structure in addition to the initial and final structures. Furthermore, we demonstrate that the characteristic dsDNA extension rates due to strand exchange and free RecA binding are the same, suggesting that relocation of the complementary strand from its position in the intermediate structure to its position in the final structure limits both rates. Finally, we propose that homology recognition is governed by transitions to and from the intermediate structure, where the transitions depend on differential extension in the dsDNA. This differential extension drives strand exchange forward for homologs and increases the free energy penalty for strand exchange of non-homologs.
E. Feinstein, C. Danilowicz, and M. Prentiss. 2012. “Mechanical Stress on Double Stranded DNA Drives Homology Recognition.” BIOPHYSICAL JOURNAL, 102, Pp. 16A-16A. Publisher's Version
J. Fisher, A. Bourniquel, G. Witz, M. Prentiss, and N. E. Kleckner. 2012. “Organization and Dynamics of the Living E. Coli Nucleoid at High Resolution in Space and Time.” BIOPHYSICAL JOURNAL, 102, Pp. 16A-16A. Publisher's Version
C. Danilowicz, E. Feinstein, A. Conover, V.W. Coljee, J. Vlassakis, Y. L. Chan, D. K. Bishop, and M. Prentiss. 2012. “RecA homology search is promoted by mechanical stress along the scanned duplex DNA.” NUCLEIC ACIDS RESEARCH, 40, Pp. 1717-1727. Publisher's VersionAbstract
A RecA-single-stranded DNA (RecA-ssDNA) filament searches a genome for sequence homology by rapidly binding and unbinding double-stranded DNA (dsDNA) until homology is found. We demonstrate that pulling on the opposite termini (3' and 5') of one of the two DNA strands in a dsDNA molecule stabilizes the normally unstable binding of that dsDNA to non-homologous RecA-ssDNA filaments, whereas pulling on the two 3', the two 5', or all four termini does not. We propose that the 'outgoing' strand in the dsDNA is extended by strong DNA-protein contacts, whereas the 'complementary' strand is extended by the tension on the base pairs that connect the 'complementary' strand to the 'outgoing' strand. The stress resulting from different levels of tension on its constitutive strands causes rapid dsDNA unbinding unless sufficient homology is present.
2011
A. J. Conover, C. Danilowicz, R. Gunaratne, V.W. Coljee, N. Kleckner, and M. Prentiss. 2011. “Changes in the tension in dsDNA alter the conformation of RecA bound to dsDNA-RecA filaments.” NUCLEIC ACIDS RESEARCH, 39, Pp. 8833-8843. Publisher's VersionAbstract
The RecA protein is an ATPase that mediates recombination via strand exchange. In strand exchange a single-stranded DNA (ssDNA) bound to RecA binding site I in a RecA/ssDNA filament pairs with one strand of a double-stranded DNA (dsDNA) and forms heteroduplex dsDNA in site I if homology is encountered. Long sequences are exchanged in a dynamic process in which initially unbound dsDNA binds to the leading end of a RecA/ssDNA filament, while heteroduplex dsDNA unbinds from the lagging end via ATP hydrolysis. ATP hydrolysis is required to convert the active RecA conformation, which cannot unbind, to the inactive conformation, which can unbind. If dsDNA extension due to RecA binding increases the dsDNA tension, then RecA unbinding must decrease tension. We show that in the presence of ATP hydrolysis decreases in tension induce decreases in length whereas in the absence of hydrolysis, changes in tension have no systematic effect. These results suggest that decreases in force enhance dissociation by promoting transitions from the active to the inactive RecA conformation. In contrast, increases in tension reduce dissociation. Thus, the changes in tension inherent to strand exchange may couple with ATP hydrolysis to increase the directionality and stringency of strand exchange.
J. K. Fisher, A. Bourniquel, M. Prentiss, and N. Kleckner. 2011. “Segregation of Sister Chromosomes in E. coli is Governed by the Shape of the Nucleoid and the Release of Inter-Sister "snaps".” BIOPHYSICAL JOURNAL, 100, Pp. 239-239. Publisher's Version
E. Feinstein, C. Danilowicz, A. Conover, R. Gunaratne, N. Kleckner, and M. Prentiss. 2011. “Single-molecule studies of the stringency factors and rates governing the polymerization of RecA on double-stranded DNA.” NUCLEIC ACIDS RESEARCH, 39, Pp. 3781-3791. Publisher's VersionAbstract
RecA is a key protein in homologous recombination. During recombination, one single-stranded DNA (ssDNA) bound to site I in RecA exchanges Watson-Crick pairing with a sequence-matched ssDNA that was part of a double-stranded DNA molecule (dsDNA) bound to site II in RecA. After strand exchange, heteroduplex dsDNA is bound to site I. In vivo, direct polymerization of RecA on dsDNA through site I does not occur, though it does in vitro. The mechanisms underlying the difference have been unclear. We use single-molecule experiments to decouple the two steps involved in polymerization: nucleation and elongation. We find that elongation is governed by a fundamental clock that is insensitive to force and RecA concentration from 0.2 and 6 mu M, though rates depend on ionic conditions. Thus, we can probe nucleation site stability by creating nucleation sites at high force and then measuring elongation as a function of applied force. We find that in the presence of ATP hydrolysis a minimum force is required for polymerization. The minimum force decreases with increasing RecA or ATP concentrations. We propose that force reduces the off-rate for nucleation site binding and that nucleation site stability is the stringency factor that prevents in vivo polymerization.
2010
E. J. Su, S. J. Wu, and M. G. Prentiss. 2010. “Atom interferometry using wave packets with constant spatial displacements.” PHYSICAL REVIEW A, 81. Publisher's VersionAbstract
A standing-wave light-pulse sequence is demonstrated that places atoms into a superposition of wave packets with precisely controlled displacements that remain constant for times as long as 1 s. The separated wave packets are subsequently recombined, resulting in atom interference patterns that probe energy differences of approximate to 10(-34) J and can provide acceleration measurements that are insensitive to platform vibrations.
A. A. Bourniquel, B. T. Ho, M. Prentiss, and N. E. Kleckner. 2010. “The E.coli Chromosome is an Internally-Organized, Springy, Helical Ellipsoid, the Shape and Dynamics of Which, Through the Cell Cycle, are Determined by the Mechanical Constraints Associated with Replication-Driven Extrusion of DNA/chromatin into a Confin.” BIOPHYSICAL JOURNAL, 98, Pp. 658A-658A. Publisher's Version
J. K. Fisher, R. Koszul, M. Prentiss, and N. Kleckner. 2010. “A Magnetic Force Micropiston for Analysis of Chromosome Expansive and Compressive Forces and their Effects on Structure and Function.” BIOPHYSICAL JOURNAL, 98, Pp. 477A-477A. Publisher's Version
J. Vlassakis, S. Tyle, T. Crawford, J. Williams, J. Weeks, T. Kodger, E. Feinstein, C. Danilowicz, V. Coljee, and M. Prentiss. 2010. “An optical tweezers study of nanosecond duration DNA conformations through DNA-surface binding distance measurements.” In OPTICAL TRAPPING AND OPTICAL MICROMANIPULATION VII, edited by K. Dholakia and G. C. Spalding. Vol. 7762. Publisher's VersionAbstract
Optical tweezers have been widely used to study DNA properties including time dependent changes in conformation; however, such studies have emphasized direct fluorescent observation of the conformations of dyed DNA molecules. In this work we explore DNA conformations that allow undyed DNA to link to spatially separated surfaces. In one set of experiments, we used optical tweezers to hold a polystyrene bead at a fixed distance from the sample capillary wall and measured the probability of the binding as a function of the separation between the polystyrene bead and the capillary, where the beads were fully confined in liquid. In a separate magnetic crystal experiment, we used magnetic forces to control the separation between magnetic beads in a hexagonal lattice at an air-water interface and measured the probability of linking to beads in the crystal. In both types of experiments peak binding occurs at a surface separation several times longer than the radius of gyration of the DNA. These experiments provide fundamental information on elusive, but significant DNA conformations, as well as technologically useful information on the probability of the DNA binding that will link two surfaces.
A. Tonyushkin and M. Prentiss. 2010. “Straight macroscopic magnetic guide for cold atom interferometer.” JOURNAL OF APPLIED PHYSICS, 108. Publisher's VersionAbstract
We demonstrate a macroscopic magnetic guide for cold atom interferometry, where the magnetic guiding field is generated by a symmetrical array of racetrack coils of copper tape. This system represents a conceptual advance over previous guided atom interferometers based on nonsymmetrical geometries because the symmetry provides a much lower magnetic field curvature per fixed length than equivalent nonsymmetrical geometries, permitting a decrease in system length without increasing the decoherence rate associated with field curvature. We realized a magnetic guide a few cm away from each coil, where smooth translation of the guided atoms is achieved by changing the currents in second array of the multiple-conductor tape. (c) 2010 American Institute of Physics. [doi:10.1063/1.3506685]
C. Danilowicz, K. Hatch, A. Conover, T. Ducas, R. Gunaratne, V. Coljee, and M. Prentiss. 2010. “Study of force induced melting of dsDNA as a function of length and conformation.” JOURNAL OF PHYSICS-CONDENSED MATTER, 22. Publisher's VersionAbstract
We measure the constant force required to melt double-stranded (ds) DNA as a function of length for lengths from 12 to 100 000 base pairs, where the force is applied to the 3'3' or 5'5' ends of the dsDNA. Molecules with 32 base pairs or fewer melt before overstretching. For these short molecules, the melting force is independent of the ends to which the force is applied and the shear force as a function of length is well described by de Gennes theory with a de Gennes length of less than 10 bp. Molecules with lengths of 500 base pairs or more overstretch before melting. For these long molecules, the melting force depends on the ends to which the force is applied. The melting force as a function of length increases even when the length exceeds 1000 bp, where the length dependence is inconsistent with de Gennes theory. Finally, we expand de Gennes melting theory to 3'5' pulling and compare the predictions with experimental results.
E. Feinstein, P. Striehl, J. Shivers, and M. Prentiss. 2010. “Using Magnetic Fluids as a Versatile Method for Manipulating and Sorting Unlabeled Nonmagnetic Particles in a Flow.” BIOPHYSICAL JOURNAL, 98, Pp. 605A-605A. Publisher's Version
2009
A. Tonyushkin, S. Wu, and M. Prentiss. 2009. “Demonstration of a multipulse interferometer for quantum kicked-rotor studies.” PHYSICAL REVIEW A, 79. Publisher's VersionAbstract
We implemented a multipulse interferometer scheme that allows us to study a quantum kicked rotor by observing dephasing of momentum coherence. Our study shows that momentum coherence can be nearly perfectly preserved under conditions where the mean energy as a function of the kick number is known to increase without bound. The accompanying width narrowing of these coherences may provide a new method for accurate measurement of the recoil frequency.
C. B. Roland, K. A. Hatch, M. Prentiss, and E. I. Shakhnovich. 2009. “DNA unzipping phase diagram calculated via replica theory.” PHYSICAL REVIEW E, 79. Publisher's VersionAbstract
We show how single-molecule unzipping experiments can provide strong evidence that the zero-force melting transition of long molecules of natural dsDNA should be classified as a phase transition of the higher-order type (continuous). Toward this end, we study a statistical-mechanics model for the fluctuating structure of a long molecule of dsDNA, and compute the equilibrium phase diagram for the experiment in which the molecule is unzipped under applied force. We consider a perfect-matching dsDNA model, in which the loops are volume-excluding chains with arbitrary loop exponent c. We include stacking interactions, hydrogen bonds, and main-chain entropy. We include sequence heterogeneity at the level of random sequences; in particular, there is no correlation in the base-pairing (bp) energy from one sequence position to the next. We present heuristic arguments to demonstrate that the low-temperature macrostate does not exhibit degenerate ergodicity breaking. We use this claim to understand the results of our replica-theoretic calculation of the equilibrium properties of the system. As a function of temperature, we obtain the minimal force at which the molecule separates completely. This critical-force curve is a line in the temperature-force phase diagram that marks the regions where the molecule exists primarily as a double helix versus the region where the molecule exists as two separate strands. We compare our random-sequence model to magnetic tweezer experiments performed on the 48 502 bp genome of bacteriophage lambda. We find good agreement with the experimental data, which is restricted to temperatures between 24 and 50 degrees C. At higher temperatures, the critical-force curve of our random-sequence model is very different for that of the homogeneous-sequence version of our model. For both sequence models, the critical force falls to zero at the melting temperature T-c like parallel to T-T-c parallel to(alpha). For the homogeneous-sequence model, alpha=1/2 almost exactly, while for the random-sequence model, alpha approximate to 0.9. Importantly, the shape of the critical-force curve is connected, via our theory, to the manner in which the helix fraction falls to zero at T-c. The helix fraction is the property that is used to classify the melting transition as a type of phase transition. In our calculation, the shape of the critical-force curve holds strong evidence that the zero-force melting transition of long natural dsDNA should be classified as a higher-order (continuous) phase transition. Specifically, the order is 3rd or greater.

Pages