Structure

Article Real-Time Observation of Strand Exchange Reaction with High Spatiotemporal Resolution Kaushik Ragunathan,1 Chirlmin Joo,2,5 and Taekjip Ha1,2,3,4,* 1Center

for Biophysics and Computational Biology of Physics, Center for the Physics of Living Cells 3Institute for Genomic Biology University of Illinois at Urbana-Champaign, Urbana, IL 61801, USA 4Howard Hughes Medical Institute, Urbana, IL 61801, USA 5Present address: Kavli Institute of NanoScience and Department of BioNanoScience, Delft University of Technology, 2628 CJ Delft, The Netherlands *Correspondence: [email protected] DOI 10.1016/j.str.2011.06.009 2Department

SUMMARY

RecA binds to single-stranded (ss) DNA to form a helical filament that catalyzes strand exchange with a homologous double-stranded (ds) DNA. The study of strand exchange in ensemble assays is limited by the diffusion limited homology search process, which masks the subsequent strand exchange reaction. We developed a single-molecule fluorescence assay with a few base-pair and millisecond resolution that can separate initial docking from the subsequent propagation of joint molecule formation. Our data suggest that propagation occurs in 3 bp increments with destabilization of the incoming dsDNA and concomitant pairing with the reference ssDNA. Unexpectedly, we discovered the formation of a dynamic complex between RecA and the displaced DNA that remains bound transiently after joint molecule formation. This finding could have important implications for the irreversibility of strand exchange. Our model for strand exchange links structural models of RecA to its catalytic function.

INTRODUCTION Double-strand breaks in DNA are catastrophic events that frequently occur during replication of lesion-containing DNA, and are principally repaired through the homologous recombination pathway (Cox et al., 2000; Kowalczykowski, 2000; Kowalczykowski et al., 1994; Spies and Kowalczykowski, 2005). During the initial stages of homologous recombination, one strand of a blunt duplex end is processed to generate a long stretch of single-stranded DNA (ssDNA). Then, in a process known as strand exchange, the ssDNA finds a homologous doublestranded DNA (dsDNA) partner in the cell and exchanges complementary base pairs to form a new heteroduplex product. Strand exchange is catalyzed by recombinases; RecA in bacteria and Rad51 in eukaryotes (Bianco et al., 1998; Sung et al., 2003). Strand exchange is followed by the formation of

Holliday junction intermediates (Potter and Dressler, 1976), which are eventually resolved by branch migration proteins to complete homologous recombination. The overall process of RecA-mediated strand exchange involves presynapsis, synapsis, and heteroduplex extension via branch migration (Bianco et al., 1998). Presynapsis involves the assembly of RecA monomers on ssDNA to form a helical filament in the presence of ATP. RecA monomers assemble on ssDNA with a stoichiometry of 3 nt per RecA monomer (Di Capua et al., 1982; Dombroski et al., 1983) and the resulting filament stretches the DNA to 1.5 times the length of B-form DNA (Dunn et al., 1982; Stasiak et al., 1981). These observations were recently confirmed by crystallographic studies of the RecA filament formed on single- and double-stranded DNA (Chen et al., 2008). Surprisingly, the DNA bound by RecA was shown to be nonuniformly stretched with triplets of nucleotides in B-form configuration and the extended conformation of the DNA arising from the large gaps between adjacent triplets (Chen et al., 2008). Single-molecule kinetic analysis revealed that RecA presynaptic filament formation is nucleated by the simultaneous binding of four to five monomers (Galletto et al., 2006; Joo et al., 2006) followed by rapid extension via monomer addition (Joo et al., 2006). After presynaptic filament formation on a ssDNA (‘‘reference’’ ssDNA), synapsis follows and this process involves at least three distinct steps (Bianco et al., 1998; Kowalczykowski, 2008). We introduce the terms of initiation, propagation and completion to describe the three steps that occur during synapsis. (1) Initiation: search for homology by the RecA filament and homologous alignment between the ‘‘incoming’’ dsDNA and the RecA-bound reference ssDNA. (2) Propagation: base-pair exchange between the reference ssDNA and incoming dsDNA molecule to form a ‘‘joint molecule.’’ Joint molecules represent a protein bound, three-stranded intermediate state during strand exchange wherein base-pair exchange may not have proceeded to completion (Menetski et al., 1990). (3) Completion: release of the ‘‘outgoing’’ displaced ssDNA from the postsynaptic complex resulting in a RecA-bound heteroduplex and free ssDNA. Joint molecule formation can be carried out efficiently without hydrolyzing ATP (Kowalczykowski and Krupp, 1995; Menetski et al., 1990), but the release of RecA from the heteroduplex product requires ATP hydrolysis (Rosselli and Stasiak, 1990).

1064 Structure 19, 1064–1073, August 10, 2011 ª2011 Elsevier Ltd All rights reserved

Structure strand exchange mechanism via single-molecule fret

The process of homology recognition by the RecA filament is a central feature of the synapsis reaction mediated by RecA. Previous studies based on chemical crosslinking proposed that the recognition process occurs via Watson-Crick pairing (Adzuma, 1992; Zhou and Adzuma, 1997). From a structural perspective, the RecA filament presents triplets of nucleotides in the B-form configuration (Chen et al., 2008) raising the intriguing possibility that base-pairing exchange between two DNA strands may proceed via Watson-Crick pairing involving destabilization of the incoming dsDNA in 3 bp increments. However, until now, there has not been an experimental test of this prediction. Following the exchange of base pairs between homologous DNA strands, the outgoing ssDNA is thought to remain bound to the RecA filament via weak interactions with the RecA secondary binding site (Mazin and Kowalczykowski, 1996, 1998). Biochemical studies demonstrated that the RecA secondary binding site serves as a gateway for strand exchange mediating the exit and the entry of DNA strands from the RecA filament (Kurumizaka et al., 1996). There is presently little information on the characteristics of DNA bound to the secondary binding site and no clear consensus exists on the structural and dynamic properties of the complex formed by the three DNA strands and the RecA filament during and after strand exchange (Camerini-Otero and Hsieh, 1993; Chiu et al., 1993; Folta-Stogniew et al., 2004; Jain et al., 1995; Podyminogin et al., 1995; Zhou and Adzuma, 1997). While fluorescence and FRET (fluorescence resonance energy transfer) based ensemble measurements have been valuable in establishing the presence of multiple kinetic intermediates during strand exchange, the number and identity of each of these intermediates remains ambiguous (Bazemore et al., 1997; Folta-Stogniew et al., 2004; Lee et al., 2006; Xiao and Singleton, 2002). Single-molecule methods are ideal for dissecting complex multistep processes by overcoming ensemble averaging and revealing reaction intermediates in real time (Weiss, 1999). RecA and its homolog Rad51 have been extensively studied using single-molecule mechanical manipulation and fluorescence based approaches (Arata et al., 2009; Hegner et al., 1999; Hilario et al., 2009; Prasad et al., 2006; Shivashankar et al., 1999; van Loenhout et al., 2009; van Mameren et al., 2009). More recently, single-molecule mechanical manipulation methods (Fulconis et al., 2006; van der Heijden et al., 2008) measured strand exchange and probed the structure of the three stranded complex, and a fluorescence based assay detected the formation of transient complexes between a ssDNA filament and a nonhomologous dsDNA (Mani et al., 2010). Here, we studied the mechanism of RecA-mediated joint molecule formation using single-molecule FRET (Ha et al., 1996). In contrast to ensemble measurements, our single-molecule FRET assay can separate the initial docking from the subsequent propagation leading to joint molecule formation thereby enabling us to analyze the strand exchange kinetics with unprecedented clarity and precision. We found that the initiation of joint molecule formation involves a synaptic complex of <14 bp in length. Our data suggest that the propagation of base-pairing leading to joint molecule formation occurs in 3 bp increments with destabilization of the incoming dsDNA and concomitant pairing with the reference ssDNA. Unexpectedly, we discovered

the formation of a highly dynamic complex between RecA and the displaced outgoing ssDNA, which remained bound for a few seconds after base-pair exchange was completed. RESULTS Single-Molecule Fluorescence Assay for Strand Exchange A biotinylated dsDNA (18 bp) with a free 50 ssDNA overhang and an acceptor fluorophore (Cy5) at the ssDNA-dsDNA junction was immobilized on a polymer-passivated surface (Figure 1A). The single-stranded portion of a specified homology length, Lh (nt), is bound by RecA to form a stable presynaptic filament using ATPgS as a cofactor. Using ATPgS allowed us to monitor synaptic events without turnover of RecA monomers from the DNA during or after reaction completion. We then flowed in a solution containing donor (Cy3)-labeled homologous dsDNA (also of homology length, Lh (bp)) and ATPgS, while simultaneously washing away free RecA in solution. This procedure ensures that the incoming dsDNA interacts solely with the immobilized filament. The labeling sites on the incoming dsDNA were chosen so that the donor and acceptor fluorophores are in close proximity after joint molecule formation. This docking-and-pairing assay monitors docking of the incoming dsDNA to the RecA filament via fluorescence signal appearance and pairing via FRET change (Figure 1A). The completion of joint molecule formation was confirmed by the appearance of a high FRET population with an apparent FRET efficiency E of approximately 0.85 (Figure 1B). A control with nonhomologous DNA produced only a low FRET population at E0.1 (Figure 1B) and the number of dsDNA bound to the RecA filaments showed a dramatic reduction (Figure 1C). In addition, the homologous dsDNA case showed a rapid accumulation of reaction products (Figure 1D) in contrast to the nonhomologous control thus recapitulating the specificity of the RecA strand exchange reaction. In order to verify that the final product formed in the presence of homologous dsDNA was the expected heteroduplex, we carried out deproteinization of the joint molecules (see Experimental Procedures) and incubation with a restriction enzyme whose restriction site was located between the donor and acceptor dyes in the final product. Over 95% of the reaction product could be cleaved off (see Figure S1 available online), resulting in the loss of donor signal and confirming that the end product is the expected heteroduplex. Direct Observation of Initial Pairing and Strand Exchange Real-time single-molecule time traces showed the docking of Cy3-labeled homologous dsDNA to the RecA filament as an abrupt appearance of fluorescence signal (Figure 2A). One class of molecules showed a low FRET value (E0.1) at the moment of docking and later transitioned to the high FRET state (E0.85) (Figure 2A, top panel). This low to high FRET transition signals successful joint molecule formation near the labeled end of DNA. The other class of molecules showed the high FRET state from the moment of docking (Figure 2A, bottom panel), indicating that joint molecule formation initiated near the labeled end. Several controls showed that the low FRET state is not due to photophysical effects of the fluorophores (Figures S2A–S2C).

Structure 19, 1064–1073, August 10, 2011 ª2011 Elsevier Ltd All rights reserved 1065

Structure strand exchange mechanism via single-molecule fret

A

B homologous dsDNA, Lh(bp) 5’

Number of Molecules

RecA 5’ Lh(nt) partial artial duplex uplex PEG EG

biotin neutravidin

A γS ATP γ ATPγS

ATPγS A ATP γ γS

800 Homologous DNA Non-homologous DNA

600 400 200 0

quartz slide l d surface f

0.0

0.2

C

0.4

0.6

0.8

1.0

FRET (E)

D

Donor

Acceptor

Homologous DNA

Donor

Number of Molecules

500

Acceptor

Non-homologous DNA

Homologous DNA Non-homologous DNA

400 300 200 100 0 0

50

100

150

200

Time (sec) Figure 1. Single-Molecule FRET Assay for Strand Exchange (A) A schematic of the FRET-based strand exchange assay. The final heteroduplex product would have the donor (green) and the acceptor (red) in close proximity. (B) FRET efficiency histograms with immobilized homologous and nonhomologous DNA (Lh = 39 nt) obtained after 10 min reaction. dsDNA of length 39 bp was used for each measurement. Data were obtained from 15 imaging areas each. See also Figure S1. (C) Images of single molecules in the donor and acceptor emission channels. The green and red circles show an example of donor and acceptor spots. Scale bar = 5 mm. (D) Number of Cy3-labeled dsDNA molecules (Lh = 39 bp) bound to the RecA filament (Lh = 39 nt) per imaging area versus time. dsDNA was added at t = 0.

We defined the term < Dtdelay > to represent the mean dwell time of the low FRET delay period (Dtdelay) prior to the final high FRET transition. We exclude the molecules that directly display high FRET upon binding while calculating the < Dtdelay > . The low FRET period cannot be attributed to the homology search process because the mean dwell time, < Dtdelay > , among those events showing nonzero Dtdelay, did not change appreciably when the ssDNA length was varied while keeping the homology length, Lh at 39 nt (Figure S2D). Therefore, homology search after docking must be instantaneous within our time resolution (30 ms) so that the initial synaptic complex formation essentially coincides with docking. It is notable that the Dtdelay distributions showed a clear shift to longer times with an increase in Lh (Figure 2B). The average of Dtdelay, < Dtdelay > , also showed a strong dependence on Lh (Figure 2C). Therefore, we attribute the initial low FRET period (Dtdelay) to the time it takes to propagate base-pair exchange from the initial synaptic complex to the labeled end of DNA (see diagrams in Figure 2A). Consistent with this interpretation, the fraction of molecules that exhibit zero delay (Figure 2A, bottom panel) decreased with increasing Lh (Figure 2C, inset). This observation is likely due to more initiation sites being available thus decreasing the probability to initiate the reaction from the labeled end. Furthermore, the histogram of the time delay (Dtdelay) exhibited by molecules showing the initial low FRET state displayed a nonexponential distribution with a peak at approximately 120 ms for the homology length (Lh) of 39 nt (Figure 2D), which we analyze in more detail in the following section.

Evidence for 3 nt Step Size of Base-Pair Exchange Having established that Dtdelay represents the propagation of joint molecule formation, we further analyzed the distribution of Dtdelay to extract information regarding the number and identity of kinetic intermediates prior to reaction completion. If there was only one rate-limiting step that needed to be overcome to complete the exchange of all the base pairs in the incoming dsDNA, we would obtain a distribution of delay times described by a single exponential curve. However, the distribution of delay times in our measurements is nonexponential and the data display a rise phase followed by a decay (Figure 2D). If the time for joint molecule formation involved N hidden rate-limiting steps prior to reaction completion, we can recapitulate the key features of our dwell time distribution plot. This approach to modeling dwell time distributions has been useful in estimating the kinetic step size of motor proteins in several biophysical studies (Myong et al., 2007; Park et al., 2010; Yildiz et al., 2003, 2004). Hence, assuming that joint molecule formation involves base-pairing exchange in well-defined increments, the distribution of Dtdelay can be used to estimate how many base pairs are exchanged per rate-limiting step: N steps

zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{ k k k k begin / /::::::/ / end: In this model, the histogram of Dtdelay should follow a gamma N1 kDtdelay distribution, Dtdelay e : N (the number of steps) and k (the

1066 Structure 19, 1064–1073, August 10, 2011 ª2011 Elsevier Ltd All rights reserved

Structure strand exchange mechanism via single-molecule fret

tdelay (sec)

300

3.5 3.0

60

40

2.5 2.0 1.5

1.0 0.8 0.6 0.4 0.2 0.0

20

40

60

1.0

20

0.0

0

100

0

1

5

0

10

15 20

0

30

10

20

30

40

E

Lh = 39 nt N = 7.7±0.4 steps k = 68±4 s-1

10

5

0.0 13

14

15

16

17

18

19

0.2

0.8

1.2

80

90

8 6

40 0

30

40

Lh (nt)

50

4 x-intercept 13.5±0.4 nt

1/slope= 3.2±0.07 bp/step

1.6

tdelay (sec)

20

70

80

2

0 12

N (Number of steps)

400 300 200 100 0

10

15

60

120

20

high FRET

11

12

Stepping rate (sec-1)

D

50

Lh (nt)

tdelay (sec)

10

80

Lh (nt)

0.5

200

high FRET

Fluorescence (a.u.)

80

Number of molecules

Fluorescence (a.u.)

high FRET

80 nt 60 nt 39 nt

Direct high FRET fraction

low FRET

C

100

< tdelay> (sec)

B Fraction of molecules

A

0 0

10

20

30

40

50

Lh (nt)

Time (sec)

Figure 2. Rapid Formation of Initial Synaptic Complex and Its Propagation in 3 bp Increments (A) Single-molecule time traces of donor (green) and acceptor (red) intensities showing docking and pairing for Lh = 39 nt. Homologous dsDNA, Lh = 39 bp was used in this measurement. In one class of events, a low FRET period of Dtdelay precedes the appearance of high FRET (top panel). In the other class of events, high FRET appears from the moment of docking (bottom panel). The cartoons above the time traces show the proposed reaction stages. The arrow in the bottom panel marks acceptor photobleaching. See also Figure S2. (B) Survival probability of the initial low FRET state versus time for Lh = 39, 60, 80 nt. dsDNA of corresponding lengths, Lh (bp), was used in each measurement. (C) < Dtdelay > versus Lh. A Power function fit was used as a guide. The population of Dtdelay = 0 was not included when calculating the average. Inset shows fraction of molecules with Dtdelay = 0 versus Lh. Exponential fitting was used as a guide. Error bars are standard errors of the mean determined from three independent experiments. (D) Gamma distribution fit of Dtdelay histograms for Lh = 39 nt. dsDNA of corresponding length, Lh = 39 bp, was used in this measurement. (E) Number of steps, N versus Lh. A linear fit of the data is shown. Inset shows the stepping rate k versus Lh. Error bars are standard errors of the mean determined from three independent experiments.

reaction rate per step) are free parameters obtained after fitting the dwell time histogram with a gamma distribution for each DNA length. From fitting the data for Lh of 31, 39, and 45 nt (Figure 2D; Figures S2E and S2F), we found that N increases with increasing Lh while k does not change significantly (Figure 2E, inset). The slope of the linear fit of N versus Lh, gave approximately 1/3 (step/bp) (Figure 2D) suggesting that base-pairing occurs in a stepwise manner in 3 bp increments. The nonzero x-intercept (Figure 2F), approximately 14 bp, specifies the number of base pairs already exchanged for N = 0, providing an estimate for the upper limit of the initial synaptic complex size. It is important to note that while Figure 2D shows the distribution of the delay time for Lh = 39 nt to be peaked at about 100– 120 ms, there are more data points at longer times outside of the major peak which are responsible for the inflated averaged delay time that we plot in Figure 2C. Our step size analysis was restricted to the major peak assuming that molecules outside of this distribution may arise from a kinetically distinct species. Data from the DNA molecules with larger Lh were not analyzed in the same manner because they exhibited broad distributions

with long time tails possibly due to multiple initiation sites along the DNA (Figure 2B). No Coupling between RecA Filament Dissociation and Joint Molecule Formation In order to ensure that the kinetics of joint molecule formation is not influenced by RecA turnover from the DNA, we carried out identical measurements using ATP as a cofactor. The same < Dtdelay > delay was observed when ATP was used as the cofactor (Figures S3A and S3B) suggesting that the rate of joint molecule formation measured is independent of the cofactor used. To test if ATP hydrolysis-mediated dissociation of RecA from the heteroduplex product might affect the rate of joint molecule formation, we modified the donor and acceptor positions in the docking-and-pairing assay. For this measurement, the reference ssDNA was labeled at an internal position (position 8 of Lh = 39 nt). The homologous incoming dsDNA (Lh = 39 bp) was also labeled internally such that after heteroduplex formation, the dyes are finally separated by 9 bp (Figure 3A). Upon docking, molecules exhibited the initial low FRET delay period (Dtdelay) prior to the appearance of a mid-FRET state

Structure 19, 1064–1073, August 10, 2011 ª2011 Elsevier Ltd All rights reserved 1067

Structure strand exchange mechanism via single-molecule fret

A 3‘

3‘

3‘

inte dye inter distance= 9bp dista A ATP

A ATP [Δtdelay]

RecA dissociation via ATP hydrolysis [τdissociation]

Number of molecules

C 3‘

14

10

Lh = 39 nt N = 3.9±0.3 steps k = 33±3 s-1

6

2 0.2

τdissociation 1000

0.4

0.8

1.0

1.2

1.4

Δtdelay (sec)

Δtdelay

D

500

Average FRET (E)

FRET (E)

B

Fluorescence (a.u.)

0.0

0

0.8 0.4

0.8

0.6

τdissociation= 29± 0.4s

0.5

0.4 0.0 140

150

160

170

180

0

190

50

100

150

200

Time(sec)

Time(sec)

Figure 3. ATP Hydrolysis-Mediated Dissociation of RecA from the Nascent Heteroduplex (A) A schematic of the modified FRET-based docking-and-pairing assay. Cy3-labeled dsDNA (Lh = 39 bp) binds to a RecA filament formed on an acceptor-labeled reference ssDNA (Lh = 39 nt) such that the distance between the dyes in the final product is 9 bp. (B) Single-molecule time traces of donor (green) and acceptor (red) intensities display a low FRET(approximately 0.1) upon initial binding (Dtdelay) corresponding to dsDNA docking followed by FRET change to a mid-FRET state (approximately 0.4) indicating the RecA-bound state of the heteroduplex finally followed by RecA dissociation via ATP hydrolysis (tdissociation) leading to a high FRET state (approximately 0.7). The arrow in the top panel marks donor photobleaching and loss of FRET. See also Figure S3. (C) Gamma distribution fit for the initial low FRET period prior to heteroduplex formation. This construct monitors propagation over a 17 nt segment of the DNA. (D) Averaged FRET trajectory for the time period, tdissociation and a single exponential fit for the RecA filament dissociation time.

(E0.4) corresponding to the stretched conformation of the heteroduplex product indicating that RecA still remains bound. Eventually, when the joint molecule is converted to a protein free heteroduplex due to RecA dissociation via ATP hydrolysis (tdissociation), molecules exhibited a high FRET state (E 0.7) (Figure 3B). We first analyzed the delay period between initial docking and FRET change (Dtdelay) to obtain a dwell time histogram which after gamma distribution fitting gave us a value of N = 4 steps (Figure 3C). In this reaction because the reference ssDNA is internally labeled, we are monitoring base-pair exchange only until position 8 of Lh = 39 nt. After accounting for the initial synaptic complex size of approximately 14 bp, this DNA construct would be effectively measuring propagation over a 17 nt region (i.e., 39 minus 8 minus 14). This corresponds to approximately 4 nt step size in propagation of joint molecule formation which is within 30% of the value supported by the main data of the paper obtained from constructs labeled at the terminal ends of the DNA (Figure 2E). To measure the time for RecA dissociation via ATP hydrolysis after completion of joint molecule formation, we postsynchronized the FRET trajectories followed by fitting with a single exponential curve to obtain a dissociation time (tdissociation) of approximately 30 s (Figure 3D). Hence, RecA filament dissociation from the heteroduplex via ATP hydrolysis occurs on a much slower timescale than the time for joint molecule formation.

Strand Separation and Joint Molecule Formation Are Concomitant Events To test our model further and measure the correlation between the kinetics of strand separation and joint molecule formation which was measured using the previous assay, we designed an alternative labeling strategy by attaching both fluorophores on the incoming dsDNA so that FRET reports on its local strand separation process (Figure 4A). For this measurement, we immobilized a DNA molecule of homology length, Lh = 39 nt, with no fluorescent label and formed a RecA filament with ATPgS as a cofactor. We observed DNA docking to the RecA filament as an abrupt appearance of fluorescence signal in the high FRET state, E0.85 (Figure 4B). Surprisingly, the disappearance of high FRET, an indication of strand separation (Ha et al., 2002; Myong et al., 2007), was followed by a period of rapid FRET fluctuations (marked by Dt2 in Figure 3B) which lasted 3.3 s on average (Figure 4D). Analogous to the docking-and-pairing assay, the high FRET period where the incoming dsDNA binds to the RecA filament in an intact conformation can be attributed to events where the initiation of joint molecule formation occurs at a position distal to the labeled end. The distance between the dyes is insensitive to the propagation of joint molecule formation until the reaction proceeds to the labeled end (see diagrams in Figure 4B). We measured the dwell time of the high FRET period (Dt1), which would represent the time taken for strand separation at the labeled end. The dwell time of the initial high FRET state

1068 Structure 19, 1064–1073, August 10, 2011 ª2011 Elsevier Ltd All rights reserved

Structure strand exchange mechanism via single-molecule fret

Figure 4. Strand Separation Kinetics

homologous dsDNA, Lh(bp)

A 5’

RecA

5’

ATPγS A ATP γ γS

A ATP ATPγS γ γS

Lh(nt)

high FRET

C

FRET (E) Fluorescence (a.u.)

dynamic FRET fluctuation

30

20

0.0

1.0 0.6 0.2 <Δt > <Δt

>

4.0

6.0

1 delay d Lh= 39 nt Nseparation = 7±0.5 ste steps k = 63±5 s-1

10

0

Δt2 (sec)

Time (sec)

high FRET

Number of molecules

B

0.2

0.4

2.0

Δt1 (sec)

(A) A schematic of the FRET-based strand separation assay. A double-labeled dsDNA (Lh = 39 bp) binds to a RecA filament formed on an unlabeled reference ssDNA (Lh = 39 nt). Only the donor-labeled strand remains on the surface after the reaction and the acceptor-labeled outgoing ssDNA is eventually released. (B) Single-molecule time traces of donor (green) and acceptor (red) intensities display a high FRET period upon binding (Dt1) followed by rapid fluctuations in FRET for a period Dt2 as shown until disappearance of acceptor signal. (C) Histogram of Dt1 and a fit to Gamma distribution. Inset shows that < Dt1 > and < Dtdelay > are similar. The population of Dtdelay = 0 was not included when calculating the average. Error bars are standard errors of the mean determined by bootstrapping. (D) Histogram of Dt2 and a single exponential decay fit.

the outgoing ssDNA and the heteroduplex product (Figure 5B). In addition, the lifetime of 400 the three-stranded complex was identical 40 200 within error for these two configurations (Fig0 30 ure 5C). In comparison, photobleaching time1.0 scale was about an order of magnitude longer 20 0.8 Δt2= 3.3 sec (Figure S4). Furthermore, the lifetime of the 0.6 10 0.4 three-stranded complex decreased substan0.2 tially when SSB was added (Figure 5D; Mazin 0 0.0 0 5 10 15 20 25 30 26 20 21 22 23 24 25 and Kowalczykowski, 1998). To monitor the Δt2 (sec) Time (sec) removal of the outgoing ssDNA by SSB, we formed a presynaptic filament on a reference ssDNA (Lh = 80 nt) and flowed a solution upon incoming DNA binding, Dt1, displayed a narrowly peaked containing homologous dsDNA along with SSB protein. Lh of distribution (Figure 4C), similar to the Dtdelay distribution 80 nt was used for the SSB analysis because an SSB tetramer observed in the previous docking-and-pairing assay. In addition, requires a minimum of approximately 65 nt for binding ssDNA Nseparation, the number of rate-limiting steps present during Dt1 under our buffer conditions using 10 mM Mg+2 and 100 mM obtained by a gamma distribution fit and < Dt1 > were in agree- Na+ (Lohman and Overman, 1985). ment with those obtained for Dtdelay (Figure 4C and inset). The strong correlation between the kinetic rates measured in these two assays suggests that strand separation in the incoming DISCUSSION dsDNA and joint molecule formation with the reference ssDNA proceed concomitantly. Search for Homology Homology search constitutes the first event of synapsis Rapid Motion of the Outgoing Strand wherein the ssDNA filament binds to a dsDNA and searches Rapid FRET fluctuations observed in the strand separation for homology. Previous studies were unable to detect any assay persisted long after joint molecule formation must have sliding of the filament on the dsDNA and concluded that 3D finished, that is Dt2 > > Dt1 (Figure 4A). Therefore, the outgoing random collision is a sufficient mechanism for homology search strand remains associated with the heteroduplex product (Adzuma, 1998). In our measurement, the encounter between after base-pair exchange has been completed. In order to two homologous DNA molecules coincided with the formation confirm that the rapid fluctuations in FRET arise from confor- of the initial synaptic complex and thus it places an upper limit mational changes between the outgoing ssDNA and the of 30 ms (our time resolution) as the timescale of homology nascent heteroduplex product, we used an alternative labeling search. However, the current configuration of our single-molescheme where the donor is attached to the outgoing strand and cule assay is not suitable for testing the role of diffusion during the acceptor to the reference strand immobilized on the surface the search for homology. The use of short oligonucleotide (Figure 5A). We reproduced the observation that the outgoing substrates limits our ability to measure events prior to the DNA strand remains bound to the RecA filament after base- formation of the initial synaptic complex as this event is rapidly pairing exchange. This assay also showed the extended period completed. We cannot exclude an earlier step in the reaction of FRET fluctuations which we observed in the previous strand pathway which could represent an intermediate during the separation assay implying large scale relative motion between homology search process. 600

Δt1 (sec)

D

Number of molecules

800

Structure 19, 1064–1073, August 10, 2011 ª2011 Elsevier Ltd All rights reserved 1069

Structure strand exchange mechanism via single-molecule fret

A

Figure 5. Dynamics of the Outgoing DNA Strand

homologous dsDNA, Lh(bp) 5’

ec RecA 5’ Lh(nt)

ATPγS A ATP γ γS

ATPγS A γS ATP γ

B dynamic FRET fluctuation

SSB

outgoing DNA dissociation Δt2 (sec)

50 40 30

Δt2(sec)= 2.9 sec

20 10 0 0

400

0

1.0 0.8 0.6 0.4 0.2 0.0

80 60 40 20 0 0

51

52

53

54

55

10

15

20

25

3 bp steps. This finding provides support for a model in which the homologous alignment of the incoming dsDNA with with 100nM SSB the RecA-bound ssDNA occurs in increwithout SSB ments of 3 bp and involves the local exchange of base pairs (Adzuma, 1992). Because of the gap between adjacent triplets observed in the RecA crystal structure (Chen et al., 2008), the incoming dsDNA needs to be stretched locally to 100 150 200 250 Δt2 (sec) mediate alignment of successive triplets. The gap between adjacent triplets in the RecA-DNA complex might provide an explanation for the rate-limiting step which we propose exists with the periodicity of 3 bp. The distortion upon stretching the 3 bp segment would unstack the bases at the gap and melt three base pairs. The newly freed triplet of nucleotides would then base pair with the triplet in the reference ssDNA. As the DNA becomes longer, the delay time increased nonlinearly (Figure 2D) indicating that joint molecule formation becomes substantially slower for longer homology lengths. The nonlinear dependence of the delay period is possibly due to the effect of multiple synaptic complexes and/or DNA topology. The current two-color FRET assay cannot distinguish between initiation from the end versus initiation in the middle of the filament. However, a four-color FRET measurement suggests that the two pathways of joint molecule formation may proceed with different kinetics (Lee et al., 2010). In addition, previous magnetic tweezers studies on topologically constrained DNA could not observe strand exchange without negative supercoiling and the apparent rate of strand exchange was much slower, approximately 2 bp/sec, indicating that DNA topology may pose a significant barrier to the propagation of strand exchange (van der Heijden et al., 2008). Another possibility for the slower global strand exchange rate measured is the effect of torsional stress due to the concomitant rotation of dsDNA and the ssDNARecA filament complex (Honigberg and Radding, 1988; Rosselli and Stasiak, 1990), which could result in the slower joint molecule formation rates in the case of longer homology lengths (Figure 2C). Δt2 (sec)

100

200

50

5

D Fraction of molecules

FRET (E)

Fluorescence (a.u.)

SSB + outgoing DNA

Number of molecules

C

(A) A schematic shows the binding of a donorlabeled dsDNA (Lh = 39 bp) to a RecA filament formed around an acceptor-labeled reference ssDNA (Lh = 39 nt) and the release of the donorlabeled outgoing ssDNA. (B) Single-molecule time traces of donor (green) and acceptor (red) intensities exhibit rapid fluctuations in FRET over a period marked by Dt2 until signal disappearance likely due to outgoing ssDNA release. (C) Histogram of Dt2 and a single exponential decay fit. See also Figure S4. (D) SSB (100 nM) flowed along with homologous dsDNA facilitates the removal of the outgoing DNA strand (Lh = 80 nt).

56

Time (sec)

Formation of the initial synaptic complex Our observations suggest that the formation of the initial synaptic complex is coincident with homology recognition. The upper limit to the size of the initial synaptic complex, which we estimated as <14 bp, is in good agreement with the earlier estimate of approximately 15 nt as the minimum length required for homology search and strand exchange (Hsieh et al., 1992). The size is also consistent with the upper limit of 16 bp placed on the synaptic length during homology search (van der Heijden et al., 2008). Perturbations that destabilize the incoming dsDNA accelerate the formation of the initial synaptic complex (Lee et al., 2006). As for oligonucleotide substrates, the thermal breathing of duplex ends may provide preferred initiation sites for strand exchange. Based on the presence of single-molecule time traces which can be categorized into two kinetically distinct populations (Figure 2A), we propose that the reaction primarily initiates from either the proximal or distal end relative to the fluorescent labels. Our data also show that the preference for the ends decreases for longer homology lengths (Figure 2C inset) consistent with the idea that more initiation sites become available with increasing DNA homology length. Propagation of Base-Pair Exchange The kinetic analysis of our data for Lh % 45 nt allowed us to propose a model wherein joint molecule formation occurs in

50

1070 Structure 19, 1064–1073, August 10, 2011 ª2011 Elsevier Ltd All rights reserved

Structure strand exchange mechanism via single-molecule fret

Dynamic Interactions between DNA and RecA Secondary Binding Site The secondary binding site of RecA binds to the outgoing ssDNA during strand exchange (Mazin and Kowalczykowski, 1998). The binding of the outgoing ssDNA to the secondary binding site signals the completion of strand exchange (Mazin and Kowalczykowski, 1996). Given that our assay can detect the completion of joint molecule formation (Figure 2A), we propose that the fluctuations in the outgoing ssDNA interaction that we detect after strand separation (Dt2, Figure 4B) might represent the bound state of DNA to the RecA secondary binding site. The structure of the three-stranded complex upon joint molecule formation remains ambiguous with evidence pointing to the existence of a metastable three-stranded structure (FoltaStogniew et al., 2004; Voloshin and Camerini-Otero, 2004) and an alternative model proposing that the outgoing ssDNA strand is stabilized by the RecA secondary binding site where it remains bound until dissociation (Mazin and Kowalczykowski, 1998). We can rule out the existence of a stable triplex structure since our data demonstrate that the three-stranded structure formed in the presence of RecA is highly dynamic with the outgoing ssDNA displaying large excursions in FRET. Hence, our data are best explained by a model in which, following the separation of the two strands of the incoming dsDNA, the outgoing ssDNA is relayed to the secondary binding site of the RecA filament where it remains bound until dissociation or until its removal is facilitated by SSB (Figure 5D). Because strand exchange is nearly isoenthalpic in terms of base-pairing with the reference ssDNA and the base-pairing that is lost in the incoming dsDNA, it remains unknown how strand exchange can be irreversible even in the absence of ATP hydrolysis. The dynamic nature of the outgoing ssDNA occupying the secondary site indicates its high conformational entropy even while it still remains bound to the RecA filament. We suggest that the entropy gain from the dynamic mode of interaction between the outgoing ssDNA and the RecA secondary binding site may provide a driving force for making the propagation of base-pair exchange irreversible (i.e., unidirectional) in the absence of ATP hydrolysis. Broader Implications Recent structural modeling of Rad51 with DNA proposed that the DNA bases were nonuniformly stretched with triplets being maintained in B-form configuration (Reymer et al., 2009) suggesting that the mechanism of strand exchange could be evolutionarily conserved across different recombinases. The tools and assays developed to study RecA-mediated strand exchange can be applied to the study of Rad51 filaments and its accessory protein partners. Our single-molecule approach can also be easily extended to other systems where homology or target search processes occur. For example, in eukaryotes, RNA interference is executed when small RNA-loaded RISC binds to a target mRNA guided by sequence recognition of 7–8 nt (Ameres et al., 2007; Bartel, 2009; Wang et al., 2009). It should be possible to observe the initial complex formation followed by the propagation of base-pairing, the cleavage of the target mRNA, and the release of the decay products in real time.

EXPERIMENTAL PROCEDURES Single-Molecule Strand Exchange Assay The quartz (Finkenbeiner) surface is passivated with Polyethylene glycol (m-PEG-5000; Laysan Bio Inc.) and 1%–2% biotinylated PEG (biotin-PEG5000; Laysan Bio Inc.). The coating of the quartz imaging surface with PEG (Roy et al., 2008) eliminated effects due to nonspecific binding of proteins. Acceptor-labeled reference ssDNA molecules were immobilized on the passivated surface by means of a biotin-neutravidin interaction. After washing away excess of acceptor molecules, the reference ssDNA was incubated with 1 mM RecA (New England Biolabs) and 1 mM ATPgS (Calbiochem) in an incubation buffer containing 25 mM Tris Acetate (pH 7.5), 100 mM sodium acetate and 1 mM magnesium Acetate. In some cases, 1 mM ATP was used instead of ATPgS. After incubation for 15 min to ensure complete filament formation on the reference ssDNA molecules, the buffer in the chamber was exchanged with a solution of homologous 500 pM dsDNA and 1 mM ATPgS in a strand exchange buffer (25 mM Tris Acetate (pH 7.5), 100 mM sodium acetate, 10 mM magnesium acetate) supplemented with an oxygen scavenging system (1 mg/ml glucose oxidase, 0.8% glucose, 0.04 mg/ml catalase and 3 mM Trolox). Imaging was initiated as soon as the buffer exchange was complete. All measurements were carried out at room temperature (23 ± 1 C). In addition to the above components, we also added SSB (a generous gift from Dr. T.M. Lohman, Washington University) depending on the experimental scheme. DNA sequences used in these measurements are listed in Table S1. Restriction Enzyme Assay Test for Heteroduplex Formation After joint molecule formation (Lh = 39 nt), we removed RecA from the DNA by exchanging the solution in the imaging chamber with a buffer containing no magnesium (10 mM Tris-Cl, 50 mM NaCl [pH 8.0]). The restriction enzyme DdeI (New England Biolabs) was suspended in the vendor supplied reaction buffer and flowed into the channel containing the DNA. After an incubation period of 30 min at 37 C, we counted the number of molecules that remained bound to the surface. Appropriate controls were carried out to ensure that there was no nonspecific cleavage activity under the conditions used. Single-Molecule Data Acquisition Excitation of the donor, Cy3, was carried out using a Nd:YAG laser (532 nm, 75 mW, Crystalaser) by means of prism type total internal reflection microscopy (Roy et al., 2008). After filtering the scattered excitation light using a 550 nm long pass filter, fluorescence emission from the donor and the acceptor was refocused onto an EMCCD camera (Andor). The Cy3 and Cy5 emissions were split into two channels using a 630 nm dichroic mirror. The time resolution for all single-molecule strand exchange experiments was 30 ms unless otherwise specified. Data measuring RecA dissociation in the presence of ATP were acquired at 100 ms, while the measurements of Dtdelay for the same experimental configuration were obtained using 30 ms time resolution. The data acquisitions were carried out using home-built software written in Visual C++. The movies obtained with the CCD were analyzed first using IDL and the intensities of the fluorophores, and the time traces were visualized using customized MATLAB programs (Joo and Ha, 2008; Roy et al., 2008). Single-Molecule Data Analysis The dwell time analysis was carried out by a home-written MATLAB program. The background intensity in the donor and acceptor channel was subtracted followed by leakage subtraction of the donor signal to the acceptor channel. Details regarding the acquisition and analysis are based on previously published methods. After visually inspecting the acquired data, we manually selected the relevant time intervals for analysis and used Origin 8.0 to plot the data. Software for acquiring and analyzing single-molecule FRET data is freely available for download from https://physics.illinois.edu/cplc/software. SUPPLEMENTAL INFORMATION Supplemental Information includes Supplemental Experimental Procedures, four figures, and one table and can be found with this article online at doi:10.1016/j.str.2011.06.009.

Structure 19, 1064–1073, August 10, 2011 ª2011 Elsevier Ltd All rights reserved 1071

Structure strand exchange mechanism via single-molecule fret

ACKNOWLEDGMENTS This work was supported by a National Science Foundation award (PHY0646550) to T.H. We thank K. Lee, J. Yoo, and S.H. Kim for discussions and experimental help. We thank C. Liu, H. Koh, R. Roy, A. Jain, R. Vafabaksh, R. Zhou, G. Lee, Y. Ishitsuka, P. Cornish, S. Myong, S. Doganay, X. Shi, S.Syed, J. Park, and I. Cisse for advice and comments. We thank M. Schlierf, W. Hwang, M. Spies, and A. Jain for comments on the manuscript. Received: February 28, 2011 Revised: May 23, 2011 Accepted: June 7, 2011 Published: August 9, 2011 REFERENCES Adzuma, K. (1992). Stable synapsis of homologous DNA molecules mediated by the Escherichia coli RecA protein involves local exchange of DNA strands. Genes Dev. 6, 1679–1694. Adzuma, K. (1998). No sliding during homology search by RecA protein. J. Biol. Chem. 273, 31565–31573. Ameres, S.L., Martinez, J., and Schroeder, R. (2007). Molecular basis for target RNA recognition and cleavage by human RISC. Cell 130, 101–112. Arata, H., Dupont, A., Mine´-Hattab, J., Disseau, L., Renodon-Cornie`re, A., Takahashi, M., Viovy, J.L., and Cappello, G. (2009). Direct observation of twisting steps during Rad51 polymerization on DNA. Proc. Natl. Acad. Sci. USA 106, 19239–19244. Bartel, D.P. (2009). MicroRNAs: target recognition and regulatory functions. Cell 136, 215–233. Bazemore, L.R., Folta-Stogniew, E., Takahashi, M., and Radding, C.M. (1997). RecA tests homology at both pairing and strand exchange. Proc. Natl. Acad. Sci. USA 94, 11863–11868. Bianco, P.R., Tracy, R.B., and Kowalczykowski, S.C. (1998). DNA strand exchange proteins: a biochemical and physical comparison. Front. Biosci. 3, D570–D603. Camerini-Otero, R.D., and Hsieh, P. (1993). Parallel DNA triplexes, homologous recombination, and other homology-dependent DNA interactions. Cell 73, 217–223. Chen, Z., Yang, H., and Pavletich, N.P. (2008). Mechanism of homologous recombination from the RecA-ssDNA/dsDNA structures. Nature 453, 489–484. Chiu, S.K., Rao, B.J., Story, R.M., and Radding, C.M. (1993). Interactions of three strands in joints made by RecA protein. Biochemistry 32, 13146–13155. Cox, M.M., Goodman, M.F., Kreuzer, K.N., Sherratt, D.J., Sandler, S.J., and Marians, K.J. (2000). The importance of repairing stalled replication forks. Nature 404, 37–41.

Ha, T., Enderle, T., Ogletree, D.F., Chemla, D.S., Selvin, P.R., and Weiss, S. (1996). Probing the interaction between two single molecules: fluorescence resonance energy transfer between a single donor and a single acceptor. Proc. Natl. Acad. Sci. USA 93, 6264–6268. Ha, T., Rasnik, I., Cheng, W., Babcock, H.P., Gauss, G.H., Lohman, T.M., and Chu, S. (2002). Initiation and re-initiation of DNA unwinding by the Escherichia coli Rep helicase. Nature 419, 638–641. Hegner, M., Smith, S.B., and Bustamante, C. (1999). Polymerization and mechanical properties of single RecA-DNA filaments. Proc. Natl. Acad. Sci. USA 96, 10109–10114. Hilario, J., Amitani, I., Baskin, R.J., and Kowalczykowski, S.C. (2009). Direct imaging of human Rad51 nucleoprotein dynamics on individual DNA molecules. Proc. Natl. Acad. Sci. USA 106, 361–368. Honigberg, S.M., and Radding, C.M. (1988). The mechanics of winding and unwinding helices in recombination: torsional stress associated with strand transfer promoted by RecA protein. Cell 54, 525–532. Hsieh, P., Camerini-Otero, C.S., and Camerini-Otero, R.D. (1992). The synapsis event in the homologous pairing of DNAs: RecA recognizes and pairs less than one helical repeat of DNA. Proc. Natl. Acad. Sci. USA 89, 6492–6496. Jain, S.K., Cox, M.M., and Inman, R.B. (1995). Occurrence of three-stranded DNA within a RecA protein filament. J. Biol. Chem. 270, 4943–4949. Joo, C., and Ha, T. (2008). Single molecule FRET with total internal reflection microscopy. In Single Molecule Techniques: A Laboratory Manual, P.R. Selvin and T. Ha, eds. (Cold Spring Harbor, NY: Cold Spring Harbor Laboratory), pp. 3–36. Joo, C., McKinney, S.A., Nakamura, M., Rasnik, I., Myong, S., and Ha, T. (2006). Real-time observation of RecA filament dynamics with single monomer resolution. Cell 126, 515–527. Kowalczykowski, S.C. (2000). Initiation of genetic recombination and recombination-dependent replication. Trends Biochem. Sci. 25, 156–165. Kowalczykowski, S.C. (2008). Structural biology: snapshots of DNA repair. Nature 453, 463–466. Kowalczykowski, S.C., and Krupp, R.A. (1995). DNA-strand exchange promoted by RecA protein in the absence of ATP: implications for the mechanism of energy transduction in protein-promoted nucleic acid transactions. Proc. Natl. Acad. Sci. USA 92, 3478–3482. Kowalczykowski, S.C., Dixon, D.A., Eggleston, A.K., Lauder, S.D., and Rehrauer, W.M. (1994). Biochemistry of homologous recombination in Escherichia coli. Microbiol. Rev. 58, 401–465. Kurumizaka, H., Aihara, H., Ikawa, S., Kashima, T., Bazemore, L.R., Kawasaki, K., Sarai, A., Radding, C.M., and Shibata, T. (1996). A possible role of the C-terminal domain of the RecA protein. A gateway model for double-stranded DNA binding. J. Biol. Chem. 271, 33515–33524. Lee, A.M., Xiao, J., and Singleton, S.F. (2006). Origins of sequence selectivity in homologous genetic recombination: insights from rapid kinetic probing of RecA-mediated DNA strand exchange. J. Mol. Biol. 360, 343–359.

Di Capua, E., Engel, A., Stasiak, A., and Koller, T. (1982). Characterization of complexes between recA protein and duplex DNA by electron microscopy. J. Mol. Biol. 157, 87–103.

Lee, J., Lee, S., Ragunathan, K., Joo, C., Ha, T., and Hohng, S. (2010). Singlemolecule four-color FRET. Angew. Chem. Int. Ed. Engl. 49, 9922–9925.

Dombroski, D.F., Scraba, D.G., Bradley, R.D., and Morgan, A.R. (1983). Studies of the interaction of RecA protein with DNA. Nucleic Acids Res. 11, 7487–7504.

Lohman, T.M., and Overman, L.B. (1985). Two binding modes in Escherichia coli single strand binding protein-single stranded DNA complexes. Modulation by NaCl concentration. J. Biol. Chem. 260, 3594–3603.

Dunn, K., Chrysogelos, S., and Griffith, J. (1982). Electron microscopic visualization of recA-DNA filaments: evidence for a cyclic extension of duplex DNA. Cell 28, 757–765.

Mani, A., Braslavsky, I., Arbel-Goren, R., and Stavans, J. (2010). Caught in the act: the lifetime of synaptic intermediates during the search for homology on DNA. Nucleic Acids Res. 38, 2036–2043.

Folta-Stogniew, E., O’Malley, S., Gupta, R., Anderson, K.S., and Radding, C.M. (2004). Exchange of DNA base pairs that coincides with recognition of homology promoted by E. coli RecA protein. Mol. Cell 15, 965–975.

Mazin, A.V., and Kowalczykowski, S.C. (1996). The specificity of the secondary DNA binding site of RecA protein defines its role in DNA strand exchange. Proc. Natl. Acad. Sci. USA 93, 10673–10678.

Fulconis, R., Mine, J., Bancaud, A., Dutreix, M., and Viovy, J.L. (2006). Mechanism of RecA-mediated homologous recombination revisited by single molecule nanomanipulation. EMBO J. 25, 4293–4304.

Mazin, A.V., and Kowalczykowski, S.C. (1998). The function of the secondary DNA-binding site of RecA protein during DNA strand exchange. EMBO J. 17, 1161–1168.

Galletto, R., Amitani, I., Baskin, R.J., and Kowalczykowski, S.C. (2006). Direct observation of individual RecA filaments assembling on single DNA molecules. Nature 443, 875–878.

Menetski, J.P., Bear, D.G., and Kowalczykowski, S.C. (1990). Stable DNA heteroduplex formation catalyzed by the Escherichia coli RecA protein in the absence of ATP hydrolysis. Proc. Natl. Acad. Sci. USA 87, 21–25.

1072 Structure 19, 1064–1073, August 10, 2011 ª2011 Elsevier Ltd All rights reserved

Structure strand exchange mechanism via single-molecule fret

Myong, S., Bruno, M.M., Pyle, A.M., and Ha, T. (2007). Spring-loaded mechanism of DNA unwinding by hepatitis C virus NS3 helicase. Science 317, 513–516. Park, J., Myong, S., Niedziela-Majka, A., Lee, K.S., Yu, J., Lohman, T.M., and Ha, T. (2010). PcrA helicase dismantles RecA filaments by reeling in DNA in uniform steps. Cell 142, 544–555. Podyminogin, M.A., Meyer, R.B., and Gamper, H.B. (1995). Sequence-specific covalent modification of DNA by cross-linking oligonucleotides. Catalysis by RecA and implication for the mechanism of synaptic joint formation. Biochemistry 34, 13098–13108.

Stasiak, A., Di Capua, E., and Koller, T. (1981). Elongation of duplex DNA by recA protein. J. Mol. Biol. 151, 557–564. Sung, P., Krejci, L., Van Komen, S., and Sehorn, M.G. (2003). Rad51 recombinase and recombination mediators. J. Biol. Chem. 278, 42729–42732. van der Heijden, T., Modesti, M., Hage, S., Kanaar, R., Wyman, C., and Dekker, C. (2008). Homologous recombination in real time: DNA strand exchange by RecA. Mol. Cell 30, 530–538. van Loenhout, M.T., van der Heijden, T., Kanaar, R., Wyman, C., and Dekker, C. (2009). Dynamics of RecA filaments on single-stranded DNA. Nucleic Acids Res. 37, 4089–4099.

Potter, H., and Dressler, D. (1976). On the mechanism of genetic recombination: electron microscopic observation of recombination intermediates. Proc. Natl. Acad. Sci. USA 73, 3000–3004.

van Mameren, J., Modesti, M., Kanaar, R., Wyman, C., Peterman, E.J., and Wuite, G.J. (2009). Counting RAD51 proteins disassembling from nucleoprotein filaments under tension. Nature 457, 745–748.

Prasad, T.K., Yeykal, C.C., and Greene, E.C. (2006). Visualizing the assembly of human Rad51 filaments on double-stranded DNA. J. Mol. Biol. 363, 713–728.

Voloshin, O.N., and Camerini-Otero, R.D. (2004). Synaptic complex revisited; a homologous recombinase flips and switches bases. Mol. Cell 15, 846–847.

Reymer, A., Frykholm, K., Morimatsu, K., Takahashi, M., and Norde´n, B. (2009). Structure of human Rad51 protein filament from molecular modeling and site-specific linear dichroism spectroscopy. Proc. Natl. Acad. Sci. USA 106, 13248–13253. Rosselli, W., and Stasiak, A. (1990). Energetics of RecA-mediated recombination reactions. Without ATP hydrolysis RecA can mediate polar strand exchange but is unable to recycle. J. Mol. Biol. 216, 335–352. Roy, R., Hohng, S., and Ha, T. (2008). A practical guide to single-molecule FRET. Nat. Methods 5, 507–516.

Wang, Y., Juranek, S., Li, H., Sheng, G., Wardle, G.S., Tuschl, T., and Patel, D.J. (2009). Nucleation, propagation and cleavage of target RNAs in Ago silencing complexes. Nature 461, 754–761. Weiss, S. (1999). Fluorescence spectroscopy of single biomolecules. Science 283, 1676–1683. Xiao, J., and Singleton, S.F. (2002). Elucidating a key intermediate in homologous DNA strand exchange: structural characterization of the RecA-triplestranded DNA complex using fluorescence resonance energy transfer. J. Mol. Biol. 320, 529–558. Yildiz, A., Forkey, J.N., McKinney, S.A., Ha, T., Goldman, Y.E., and Selvin, P.R. (2003). Myosin V walks hand-over-hand: single fluorophore imaging with 1.5-nm localization. Science 300, 2061–2065.

Shivashankar, G.V., Feingold, M., Krichevsky, O., and Libchaber, A. (1999). RecA polymerization on double-stranded DNA by using single-molecule manipulation: the role of ATP hydrolysis. Proc. Natl. Acad. Sci. USA 96, 7916–7921.

Yildiz, A., Tomishige, M., Vale, R.D., and Selvin, P.R. (2004). Kinesin walks hand-over-hand. Science 303, 676–678.

Spies, M., and Kowalczykowski, S.C. (2005). Homologous recombination by RecBCD and RecF pathways. In The bacterial chromosome, N.P. Higgins, ed. (Washington, D.C.: ASM Press), pp. 389–403.

Zhou, X., and Adzuma, K. (1997). DNA strand exchange mediated by the Escherichia coli RecA protein initiates in the minor groove of double-stranded DNA. Biochemistry 36, 4650–4661.

Structure 19, 1064–1073, August 10, 2011 ª2011 Elsevier Ltd All rights reserved 1073

Real-Time Observation of Strand Exchange Reaction with ... - Cell Press

Aug 10, 2011 - (D) Gamma distribution fit of Dtdelay histograms for Lh = 39 nt. dsDNA of corresponding length, Lh = 39 bp, was used in this measurement.

970KB Sizes 1 Downloads 351 Views

Recommend Documents

Real-Time Observation of Strand Exchange Reaction with ... - Cell Press
Aug 10, 2011 - more detail in the following section. Evidence for 3 nt Step Size ...... The data acquisitions were carried out using home-built software written in ...

Review - Cell Press
Jun 6, 2007 - well give clues to the overall requirements for reprogram- ming in other ... ada et al., 2006). In devising their new ...... Histone code modifications ...

Article - Cell Press
6 The Hartwell Center for Bioinformatics and Biotechnology .... type (**p < 0.005, ***p < 0.0005 for the average distance of Nestin+ versus NestinА cells from CD34+ cells). ..... Daoy cells alone (left) or Daoy cells plus 500 3 103 PHECs (right).

Article - Cell Press
Keiran S.M. Smalley,6 Alka Mahale,2 Alexey Eroshkin,1 Stuart Aaronson,2 and Ze'ev Ronai1,*. 1 Signal .... ERK on stability versus transcription of c-Jun, we moni- ...... PLUS Reagent (Invitrogene) following the manufacturer's protocol.

Wolbachia trends - Cell Press
tailed dance fly, Rhamphomyia longicauda. Anim. Behav. 59, 411–421. 3 Trivers, R.L. (1972) Parental investment and sexual selection. In Sexual Selection and ...

Binocular rivalry - Cell Press
percept, binocular co-operation gives way to competition. Perception then alternates between the two eyes' images as they rival for perceptual dominance .... Psilocybin links binocular rivalry switch rate to attention and subjective arousal levels in

Requirement of Synaptic Plasticity - Cell Press
Jun 3, 2015 - [email protected] (T.K.), [email protected] (A.T.). In Brief. Kitanishi et al. identify GluR1-dependent synaptic plasticity as a key cellular.

PDF (603 KB) - Cell Press
N-terminal acetylation, one of the most common protein modifications in humans. Introduction. Researchers have ..... site-directed mutagenesis with the primers hNAA10 T109C F: 50-C .... and had only achieved a social smile as a developmental mile- st

Realtime HTML5 Multiplayer Games with Node.js - GitHub
○When writing your game no mental model shift ... Switching between different mental models be it java or python or a C++ .... Senior Applications Developer.

Reaction of trihaloisocyanuric acids with alkynes - Arkivoc
Dec 3, 2017 - Table 2. Monohalogenation of different alkynes with TXCA. R. OAc. R'. X. R. R'. 0.34 eq. TXCA. HOAc:Ac2O (1:1), r.t. ..... Spectral characterization of the products is available in the Supplementary File. References. 1. Mendonça, G. F.

Origin of Exopolyphosphatase Processivity: Fusion of an ... - Cell Press
10, 11, and 13; PDB code: 1F0J). Shown in gray is the PDE4B2B ..... transformed into the BL21(DE3) bacterial host strain (Novagen). A single colony was ...

Neuroscience-Inspired Artificial Intelligence - Cell Press
Jul 19, 2017 - The fields of neuroscience and artificial intelligence (AI) have a long and intertwined history. In more recent times, however, communication and collaboration between the two fields has become less commonplace. In this article, we arg

Conscious intention and motor cognition - Cell Press
May 5, 2005 - Conscious intention and motor cognition. Patrick Haggard. Institute of Cognitive Neuroscience, University College London, 17 Queen Square, ...

Estimating diversification rates from phylogenetic ... - Cell Press
Oct 25, 2007 - Department of Biology, University of Missouri-St Louis, MO 63121-4499, USA. Patterns of species richness reflect the balance between speciation and extinction over the evolutionary history of life. These processes are influenced by the

Endophilin Mutations Block Clathrin-Mediated ... - Cell Press
ments of endo1 and endo2 allowed recovery of viable and fertile revertants ...... Thomas Schwarz for comments, and members of the Bellen Lab. Ikeda, K., and ...

Does the brain calculate value? - Cell Press
the driver of choice. Value-based vs. comparison-based theories of choice. How does the brain help us decide between going to a movie or the theatre; renting ...

Computational Precision of Mental Inference as Critical ... - Cell Press
In both (A) and (B) error bars indicate s.e.m., and the black curve indicates the theoretical ..... shown for each variability type for the combination of biases that best fitted the subjects' behavior. ...... SciPy: Open source scientific tools for

Computational Precision of Mental Inference as Critical ... - Cell Press
Dec 1, 2016 - This experimental framework revealed that in contrast to current views, the ...... Kaufman, M.T., and Churchland, A.K. (2013). Cognitive ...

On the Perception of Probable Things: Neural Substrates ... - Cell Press
Nov 8, 2011 - ments call for a significant shift in the way we think about the neuronal ..... in MT reflects rather different functions and mechanisms. Figure 3.

Computational Precision of Mental Inference as Critical ... - Cell Press
Dec 1, 2016 - ability beyond what can be explained by the provided evidence and are thus ...... study. Drawing few samples results in noisy posterior distribu-.

Fission and Uncoating of Synaptic Clathrin-Coated ... - Cell Press
Endophilin SH3 Domain Interactions with Dynamin and Synaptojanin Are Conserved in Lamprey .... the accumulation of free coated vesicles (arrowheads), coated pits, and ..... with 100 M PP-19 (VAPPARPAPPQRPPPPSGA) or 100 M PP-15.

Phylogenomics and Morphology of Extinct Paleognaths ... - Cell Press
Dec 15, 2016 - in Madagascar and the successful recovery of nuclear genome fragments with ... The genome-scale data yield stable time estimates irrespec-.

Experimental observation of decoherence
nomena, controlled decoherence induced by collisions with background gas ... 1: (a) Schematic illustration of a SQUID. ... (b) Proposed scheme for creating.

Reprogramming cell shape with laser nano-patterning - Journal of Cell ...
Feb 22, 2012 - network. The engagement of actin filaments in between individual integrins participates to the clustering of ... interest in imaging software controlled displacements of galvanometric ... (C) Monitoring of cell shape changes.