doi:10.1016/j.jmb.2004.06.024

J. Mol. Biol. (2004) 341, 739–751

Exploring Rare Conformational Species and Ionic Effects in DNA Holliday Junctions Using Single-molecule Spectroscopy Chirlmin Joo1, Sean A. McKinney1, David M. J. Lilley2 and Taekjip Ha1* 1 Physics Department University of Illinois Urbana-Champaign, Urbana IL 61801, USA 2

Cancer Research-UK Nucleic Acid Structure Research Group MSI/WTB Complex, The University of Dundee, Dundee DD1 5EH, UK

The four-way DNA (Holliday) junction is an essential intermediate in DNA recombination, and its dynamic characteristics are likely to be important in its cellular processing. In our previous study we observed transitions between two antiparallel stacked conformations using a single-molecule fluorescence approach. The magnesium concentrationdependent rates of transitions between stacking conformers suggested that an unstacked open structure, which is stable in the absence of metal ions, is an intermediate. Here, we sought to detect possible rare species such as open and parallel conformations and further characterized ionic effects. The hypothesized open intermediate cannot be resolved directly due to the limited time resolution and sensitivity, but our study suggests that the open form is achieved very frequently, hundreds of times per second under physiologically relevant conditions. Therefore despite being a minority species, its frequent formation raises the probability that it could become stabilized by protein binding. By contrast, we cannot detect even a transient existence of the junctions in a parallel form, and the probability of such forms with a lifetime greater than 5 ms is less than 0.01%. Stacking conformer transitions are observable in the presence of sodium or hexammine cobalt (III) ions as well as magnesium ions, but the transition rates are higher for lower valence ions at the same concentrations. This further supports the notion that electrostatic stabilization of the stacked structures dictates the interconversion rates between different structural forms. q 2004 Elsevier Ltd. All rights reserved.

*Corresponding author

Keywords: Holliday junction; branch migration; parallel conformation; ionic effect; single molecule spectroscopy

Introduction Genetic recombination is an important mechanism for the repair of damage to DNA that occurs in all cells.1 The central intermediate in this process is the four-way (Holliday) junction,2 which can be described as a four-stranded DNA structure comprising four helices joined in the middle (Figure 1(a)).3 Although much is known about its static structure, there is only limited information on the dynamic properties and their functional consequences. The dynamic characteristics of the junction are likely to be important in understanding Abbreviation used: FRET, fluorescence resonance energy transfer. E-mail address of the corresponding author: [email protected]

how different enzymes recognize and process the Holliday junction. Single-molecule fluorescence techniques offer a promising avenue into these aspects, since they can reveal the structural changes of individual molecules in real time while they are functioning, and do not require synchronization.4 Indeed, single-molecule fluorescence resonance energy transfer (FRET)5,6 has been used to reveal spontaneous dynamics of DNA structures,7,8 to deduce the structures of DNA–protein complexes,9,10 and to follow the structural changes in DNA induced by protein binding or enzymatic activities.11,12 In the absence of added metal ions, the Holliday junction adopts an open conformation where the four helices are directed to the corners of a square.13,14 On addition of divalent metal ions such as magnesium, the structure folds by the pairwise, coaxial stacking of helical arms to form the

0022-2836/$ - see front matter q 2004 Elsevier Ltd. All rights reserved.

740

Conformational Species of Holliday Junction

Figure 1. Conformational polymorphism in a four-way DNA junction studied using FRET. The four arms of the junction are named B, H, R and X sequentially around the point of strand exchange. Cy3 (green circle) and Cy5 (red circle) fluorophores are attached to the ends of two arms. Vectors are named according to the arms carrying donor and acceptor (in that order). (a) The XB vector of junction 1. In the absence of metal ions the junction takes the opensquare form. On addition of divalent metal ions, the structure folds into the stacked X-structure. There are two possible conformers that differ in the stacking partners, shown in an antiparallel conformation (iso I and iso II). (b) The positions of attachment of fluorophores are reversed in the BX vector compared to the XB vector, but should exhibit the same variation in fluorescence in response to the conformational transition between iso I and iso II. (c) XR vector. The fluorophores will only be close in the antiparallel form of iso II, in contrast to the XB and BX vectors.

antiparallel stacked X-structure in which the continuous strands of the junction are oriented in opposite directions.14,15 This structure was originally proposed on the basis of electrophoretic,14,16,17 FRET15,18 and chemical probing experiments,19 and has recently been confirmed by X-ray crystallography.20 – 22 Formation of the stacked X-structure in the presence of Mg2þ lowers the symmetry of the junction, and there are two possible conformers of this structure that differ in the choice of stacking partners (Figure 1(a) –(c)). Thus a junction with arms named B, H, R and X (sequentially around the junction, as depicted in Figure 1(a)) may stack into alternative conformers based on the stacking of arms X on R (and consequently B on H), or X

on B. These are termed the iso I and iso II conformers here.23 In principle each of the stacking conformers could exist in predominantly parallel or antiparallel forms, although earlier studies have indicated that the antiparallel structure is the major form present in solution15 and in the crystal.20,22 However, in the original proposals for the mechanism of recombination the structures were expected to be parallel,2,24 and have consequently most frequently been drawn this way in textbooks. Although not observed in bulk solution measurements,14,15 it is possible that the parallel structure exists in a minor population rare enough to evade detection in ensemble experiments. Single-molecule measurements should be able to detect even very rare conformations as long as

741

Conformational Species of Holliday Junction

they have distinct fluorescence signals and they are longer-lived than the time resolution of the experiment. The four-way junction can undergo a number of important dynamic processes. First, a junction can exchange between the two alternative stacking conformers. It is not possible to synchronize this process, and it was difficult to demonstrate in solution except by rather indirect experiments.23,25 – 27 However, we have recently used single-molecule FRET analysis to demonstrate conformer exchange in a direct manner, and measured the rates of interconversion.7 Four-way junctions can also undergo branch migration by virtue of the sequence homology, in which there is a step-wise exchange of base-pairing partners that moves the point of strand exchange.28,29 Branch migration is an important element in the recombination mechanism,2,24,30 and is accelerated in the cell by the action of specific proteins.31 – 35 Both conformer exchange and branch migration require the opening of the four-way junction to permit the required exchange of stacked helices or base-pairing. In the case of branch migration this is consistent with the observed inhibition of the process by concentrations of metal ions known to stabilize the stacked X-structure.36 Here, we have used single-molecule fluorescence spectroscopy to probe for the existence of minor species, the anticipated open intermediate of conformer transitions, and the parallel conformations of the stacked form of the junction. We find that the population of parallel species is exceedingly low, and the open species is achieved very frequently, hundreds of times per second under physiologically relevant conditions even though it is too short-lived to be detected directly.

Results Non-migratable four-way DNA junctions for single molecule FRET experiments We have studied 4H DNA junctions with four perfectly base-paired arms each of 11 bp. The sequences were chosen to disallow branch migration, and are thus termed non-migratable. We have measured FRET between donor (Cy3) and acceptor (Cy5) fluorophores attached to the ends of different helical arms to distinguish various possible junction conformations as shown schematically in Figure 1(a). Fluorescent vectors are named according to the arms carrying the donor and acceptor in that order; for example the XB vector carries Cy3 on the end of the X arm, and Cy5 on the end of the B arm. We have used a number of vectors here to report on the various possible conformations of the junction. XB Vector (Figure 1(a)). The donor– acceptor pair should be physically close (and therefore result in elevated energy transfer) only for the antiparallel

conformer of the X on R stacked form. Conversion to the alternative (X on B) stacking conformer leads to separation of the fluorophores, and thus lowered energy transfer. The vector BX (Figure 1(b)) has the fluorophores reversed with respect to XB, and should behave equivalently. XR Vector (Figure 1(c)). The fluorophores will only be close in the antiparallel form of the alternative stacking conformer with X on B stacking. BR Vector (Figure 4(a)). The fluorophores will be close in either of the parallel forms. It will thus identify parallel conformers as a group, but not distinguish between stacking conformers. The fluorophores are attached to the alternative pair of helices in the XH vector, which would therefore be expected to behave equivalently to BR. The conformational interconversions of these vectors have been analyzed using single-molecule FRET spectroscopy, whereby the junction molecules were immobilized onto a glass surface via biotin covalently attached to the end of one arm that is free from fluorophore conjugation. Conformational exchange between stacking conformers of junction 1 Ensemble experiments indicated that junction 1 adopts predominantly the antiparallel iso I stacked X-structure.14 We chose the XB vector for initial study, which should exhibit high FRET efficiency for iso I (Figure 1(a)). Figure 2(a) shows a time record of donor and acceptor fluorescence signals (ID and IA, respectively, obtained by subtracting background counts and correcting for cross-talk between the two detection channels) of a single junction 1 molecule in the presence of 10 mM Tris – HCl (pH 8.0), 30 mM Mg2þ, 50 mM Naþ. ID and IA exhibit two-state fluctuations that are clearly anti-correlated. Figure 2(b) shows the histogram of the apparent FRET efficiency Eapp, defined as Eapp ; IA =ðID þ IA ). The Eapp histogram has two peaks corresponding to the two stacking conformers with a strong bias toward the higher FRET state. We assign the Eapp ¼ 0.69 state as iso I and the Eapp ¼ 0.15 state as iso II, as indicated in Figure 1. The forward and backward stacking conformer transition rates (kI!II and kII!I) were obtained through a combination of cross-correlation analysis and the ratio between numbers of data points corresponding to each FRET state (see Materials and Methods). We obtained kI!II ¼ 9.3 s21 and kII!I ¼ 28 s21, averaged over 41 molecules. Thus junction 1 significantly favors the iso I conformer as expected.14 Metal ion dependence of conformer transitions in junction 1 In our previous study of junction 7 we discovered that the conformer transitions became faster as the Mg2þ concentration was reduced. The conformer transition of junction 1 exhibits the same dependence on the concentration of Mg2þ.

742

Conformational Species of Holliday Junction

Figure 2. Conformational dynamics between stacking conformers. (a) The interconversion of the XB vector of junction 1 between two stacking conformers (8 ms bin time). It shows a time record of donor (green) and acceptor (red) fluorescence signals of a single molecule in the presence of 30 mM Mg2þ and 50 mM Naþ. (b) Histogram of the apparent FRET efficiency Eapp from 41 molecules. The high FRET ðEapp , 0:69Þ state corresponds to iso I and the low FRET ðEapp , 0:15Þ, iso II. (c) Time record in the presence of 1 mM Mg2þ and no Naþ. The acceptor photobleached at 2.6 seconds. (d) Histogram of Eapp from ten molecules. (e) Time record in the presence of 0.1 mM Mg2þ and no Naþ. (f) Histogram of Eapp from 13 molecules. (g) Cross-correlation analysis on BX vector of junction 1. The correlation time, t, changes over different Mg2þ concentrations. t is 2.1, 21.2 and 71.3 ms in the presence of 0.5 (circles), 10 (squares) and 100 mM (triangles) Mg2þ, respectively. (h) Conformer transition rates kCT of BX, XB and XR vectors of junction 1 as a function of Mg2þ concentration. Each kCT value is the average over about ten molecules. Error bars on the graph represent the heterogeneity in kCT values between molecules.

Comparison of the FRET efficiency histograms in Figure 2(b) and (d) shows that the ratio between iso I and iso II remained constant regardless of Mg2þ concentration. This behavior was true for all concentrations of Mg2þ, as long as two FRET states were distinguishable in the histogram. Thus Mg2þ

must stabilize both stacking conformers equally by screening of electrostatic repulsion between phosphate groups around the point of strand exchange. Stacking conformer transitions require the disruption of base stacking at the point of strand exchange, and should become slower if the

743

Conformational Species of Holliday Junction

stacked structure is stabilized by elevated Mg2þ concentration. In order to study the fast dynamics under more physiologically relevant conditions, cross-correlation analysis was used to provide accurate values of the conformer transitions rate kCT, where kCT ; kI!II þ kII!I . Figure 2(g) shows how the correlation time ðt ¼ k21 CT Þ changes with different Mg2þ concentrations. The resulting kCT rates (each an average of , ten molecules) are plotted as a function of ionic conditions in Figure 2(h). Error bars represent the heterogeneity in kCT values between molecules (see Materials and Methods). As illustrated in Figure 1(a), while only iso I has high FRET, low FRET states might result from any of the other conformations including the iso II, open and parallel forms. In contrast to the XB vector, high FRET efficiency is expected for the iso II conformer of the XR vector (Figure 1(c)). If the low FRET states of the XB vectors arise only from iso II, XR vectors should show the same transition rates of junction dynamics regardless of their different labeling. In Figure 2(h), transition rates measured for XB (triangles) and XR (squares) vectors fall on a single line with deviations that are small compared to the error bars, indicating that the low FRET state of XB vectors mainly consists of iso II in the presence of high Mg2þ concentration. In addition, this indicates that there is no significant effect of fluorophores or attachment position on junction dynamics. Taken together with our previous studies of junctions 3 and 7,7 the new data on junction 1 show that the two-state FRET fluctuations observed represent stacking conformer transitions, and provide further evidence of the sensitivity of the conformation of the four-way junction to the sequence immediately at the point of strand exchange (Table 1). Transitions between open and stacked conformations of junction 1 at low Mg21 concentration In order for the conformer transitions to occur, base stacking at the exchange point has to be disrupted, suggesting that the transition trajectory should pass through the open structure that is

Table 1. Stacking conformer bias in three DNA junctions as a function of metal ions Junction 1 Ion iso I (%) iso II(%) kCT (s21) a b

50 mM Mg2þ 77.9 22.1 20.1

Junction 3a 50 mM Mg2þ 22.6 77.4 15.5

Junction 7 a

50 mM Mg2þ 48.3 51.7 11.8

From our previous study.7 In the presence of 50 mM Naþ.

1M Naþ 64.3 35.7 60.0

2 mMb [Co(NH3)6]3þ 61.3 38.7 70.0

stable in the absence of divalent ions (Figure 1(a)). The same open species should lie on the trajectory for spontaneous branch migration; however, in that case the base-pairing must also be disrupted. In the branch migration process the open structure is not therefore likely to be the transition state, but instead must be a high energy intermediate. It is therefore likely that the open structure is a common intermediate in both stacking conformer transitions and branch migration. This then raises the question of whether the open structure might be detected as a stable structure of measurable lifetime. We studied the XB vector of junction 1 as the Mg2þ concentration was reduced, seeking the observation of a new peak emerging between the two peaks in the Eapp histogram shown in Figure 2(b) and (d). However, instead of revealing a new peak, the two peaks merged into a single peak as the Mg2þ concentration was reduced (Figure 2(f)). Such behavior can be fully explained if the lifetime of the states is less than the time resolution, leading to averaging of FRET values. As an alternative way of seeking the open structure, we performed experiments at low Mg2þ concentrations, where conformer transitions are not fully resolved. Figure 3 shows the timeaveraged Eapp values measured from single molecules of the XB vector of junction 1 at Mg2þ concentrations between 1 mM and 1 mM, in the absence of Naþ (circles). Since junction 1 is biased strongly toward iso I (high FRET state), Eapp averaged over multiple transitions is expected to be high, close to the Eapp of iso I, as long as junction 1 is folded into stacked conformations. This appears to be the case at Mg2þ concentrations above 100 mM Mg2þ. By contrast, at sufficiently low Mg2þ concentrations, the junction would exist primarily in the open structure with Eapp value significantly below that of iso I, which is the case below 10 mM Mg2þ. In between these extremes, in the range 10– 100 mM Mg2þ, a clear and gradual transition in the average Eapp is observed, which we interpret as the transition between open and stacked structures. This interpretation is consistent with earlier ensemble studies37 that showed that stacked X-structures become stable at , 80 mM Mg2þ. A control experiment with the XR vector did not show significant changes in Eapp (squares in Figure 3), ruling out possible Mg2þ-dependent effects on fluorophore properties as the source for the observed transition. Parallel conformations of junction 1 are not detectable While our single-molecule data and previous ensemble data collectively indicate that conformational exchanges occur predominantly between two antiparallel stacking conformers, we explored the possibility that parallel conformations might also exist in some small proportion. We studied the BR vector of junction 1, where the fluorophores

744

Conformational Species of Holliday Junction

Figure 3. Estimating the population of the open structure of junction 1. The XB (circles) and XR (squares) vectors of junction 1 have been studied as a function of Mg2þ concentration. Eapp values averaged over ,20 molecules are plotted as a function of ion concentration. The averaged Eapp of the XB vector is close to the Eapp of iso I above 100 mM Mg2þ. Below 10 mM Mg2þ, the junction exists primarily in the open structure with Eapp value significantly below that of iso I. In between, a clear and gradual transition in the average Eapp is observed. In contrast, no significant change in Eapp is observed in the XR vector.

come into close proximity only in the parallel conformations (Figure 4(a)). Time traces for this vector (Figure 4(b) and (d)) exhibited a constant level of Eapp (, 0.3) without any transition to high Eapp values, and lacking any anticorrelation between donor and acceptor intensities (Figure 4(c) and (e)) at a time resolution of 5 ms in several different solution conditions, i.e. 1, 10 and 50 mM Mg2þ with 50 mM Naþ, 30 mM and 50 mM Mg2þ without Naþ, 50 mM and 1000 mM Naþ without Mg2þ. A quantitative limit on the existence of parallel conformations was established by examining the BR vector in the presence of 30 mM Mg2þ with no Naþ, at 5 ms time resolution. No transition to a higher FRET state was observed within a total observation time of 50 seconds measured from 23 molecules. Therefore, a stable parallel conformation with a lifetime greater than 5 ms must be exceedingly rare, with a probability lower than 0.01%. We have performed the same measurements on the XH vector, which should also be sensitive to the presence of parallel conformations. In addition, no evidence of parallel conformations was obtained for this vector in the presence of 50 mM Mg2þ at 6 ms time resolution (Figure 4(f) and (g)). In contrast to these properties of DNA junctions, we have observed a significant population (, 25%) of the parallel form of a four-way RNA junction, using equivalent labeling strategy in singlemolecule experiments.38 The failure to detect the parallel conformation of the DNA junction is therefore unlikely to be an artifact of the experimentation, and we conclude that its population is vanishingly small for four-way DNA junctions. Transitions between stacking conformers of junction 7 in the presence of monovalent ions alone If the role of Mg2þ is primarily in the screening of electrostatic interactions, high concentrations of monovalent ions may achieve the same effect as Mg2þ. Indeed, in previous studies the stacked-X

structure has been observed in the presence of monovalent ions only (e.g. 1 M Naþ).37 We have therefore used single-molecule FRET spectroscopy to examine whether the stacking conformer transitions can be observed in Naþ alone, how Naþ concentration affects the transition rates, and whether the conformer bias seen in Mg2þ is maintained. The transition rate of junction 7 is lower than that of junction 1 (Table 1), thus measurements on junction 7 are more accessible. So, junction 7 was used to examine the ionic effects of monovalent and trivalent ions. We studied the HB vector of junction 7 at high Naþ concentrations (400 mM to 2 M) in the absence of Mg2þ. The buffer included 2.5 mM or 5 mM EDTA to chelate any residual divalent ions. Figure 5(a) shows the time record of a single junction 7 molecule in the presence of 1.5 M Naþ. The corresponding Eapp histogram (Figure 5(b)) comprises two peaks, although there is broad overlap that probably arises from time-averaging due to fast exchange between states. iso I is favored over iso II by 3 to 2, in contrast to the 1:1 distribution observed in the presence of Mg2þ (Table 1). The transition rates averaged over 13 molecules were kI!II ¼ 24 s21 and kII!I ¼ 36 s21. Figure 5(e) shows transition rates for junction 7 measured as a function of Naþ concentration from 400 mM to 2 M. As in the case of Mg2þ, increased Naþ concentration in this range reduced the rate of the conformer transitions, while leaving the ratio between iso I and iso II unchanged (data not shown). Therefore, the effect of Naþ appears to be very similar to that of Mg2þ, except that much higher concentrations are needed to stabilize the stacked structures. Transitions between stacking conformers of junction 7 in the presence of hexammine cobalt (III) ions Hexammine cobalt (III) ions have a similar octahedral geometry to Mg2þ, but differ in two

Conformational Species of Holliday Junction

745

Figure 4. The parallel forms of the junction are not detected. (a) The donor – acceptor pair for the BR vector of junction 1 should come into close proximity only in the parallel forms. The XH vector (not shown) would similarly detect the parallel form. Typical single-molecule time records with 5 –6 ms time resolution are shown for the BR vector of junction 1 in 30 mM Mg2þ (b), 1 mM Mg2þ (d) and for the XH vector in 50 mM Mg2þ (f). In each case, donor and acceptor fluorescence signals do not exhibit any anti-correlated two-state fluctuations except in the case of photoblinking/photobleaching events of the acceptor. (c), (e) and (g) Cross-correlation analysis from about 30 molecules in each case, which supports non-existence of parallel forms.

important respects. The charge on the ion is þ 3, and the NH3 ligands of the cobalt ion are very good H-bond donors. These ions are highly efficient in promoting structural transitions in DNA, including stabilizing left-handed Z-DNA39 for example. Previous studies have demonstrated the formation of the stacked X-structure in micromolar concentrations of [Co(NH3)6]3þ.37 We tested whether the stacking conformer transitions can be observed in [Co(NH3)6]3þ, how the transition rates are affected by the concentration of the trivalent ion, and whether the conformer bias observed in the presence of Mg2þ and Naþ is maintained. Figure 5(c) shows a time record of FRET efficiency of junction 7 in the presence of 2 mM [Co(NH3)6]3þ, 50 mM Naþ, 0.1 mM EDTA. The corresponding Eapp histogram (Figure 5(d)) contains two peaks. The iso I conformation (lower

FRET state) is favored over iso II by 3 to 2 in contrast to the 1:1 partitioning observed in Mg2þ solution, (Table 1), consistent with previous studies using gel electrophoresis.23 The transition rates averaged over 16 molecules were kI!II ¼ 28 s21 and kII!I ¼ 42 s21. Figure 5(f) shows transition rates measured from junction 7 as a function of [Co(NH3)6]3þ concentration in the range from 0.1 mM to 2 mM in the presence of 50 mM Naþ and 0.1 mM or 0.25 mM EDTA. Increased [Co(NH3)6]3þ concentration decreased the conformational transition rates, similar to the effects of Mg2þ and Naþ. In addition, the ratio between iso I and iso II was constant over this range of [Co(NH3)6]3þ concentration, as it was with other ions (data not shown). Therefore, while the [Co(NH3)6]3þ promotes the formation of the stacked X-structure of the DNA junction at much

746

Conformational Species of Holliday Junction

Figure 5. Dynamics of stacking conformer exchange of junction 7 in the presence of Naþ and hexammine Co (III) ions. Data were obtained from the junction 7 HB vector. (a) Fluorescence signals with time resolution of 4 ms, in the presence of 1.5 M Naþ. The acceptor photobleached at 3.25 seconds. (b) The histogram of Eapp values measured from 13 molecules in the same condition as (a). Eapp fluctuates between ,0.66 (iso II) and , 0.19 (iso I). (c) Time record in the presence of 2 mM [Co(NH3)6]3þ and 50 mM Naþ. The acceptor photobleached at 2.25 seconds followed by the donor photobleaching at 2.35 seconds. (d) The histogram of Eapp from 16 molecules in the same condition as (c). (e) The conformer transition rate kCT as a function of Naþ concentration. (f) kCT as a function of [Co(NH3)6]3þ concentration in the presence of 50 mM Naþ. Each rate is an average from , ten different molecules, and error bars reflect the heterogeneity in kCT values between molecules.

lower concentrations, the effects on the dynamics of the interconversion are closely similar to those of other metal ions. Competition and cooperation between monovalent and divalent ions In an earlier study,7 we found that the rate of stacking conformer transitions occurring in Mg2þcontaining solution was accelerated by the addition of 50 mM Naþ. However, we have seen that high concentrations of Naþ as the only cation slow the rate of transitions (Figure 5(e)). Naþ thus play dual roles in determining conformational dynamics, and it is therefore expected that the interconversion rate in the presence of Mg2þ will exhibit a non-monotonic dependence on Naþ con-

centration, with a maximum rate at some concentration. Figure 6 shows the dependence of transition rates on Naþ concentration in the presence of three different background Mg2þ concentrations of 2, 5 and 20 mM. Clearly, the overall trend is non-monotonic, with a maximum rate in each case. For example, at 5 mM Mg2þ, addition of Naþ to a concentration of 300 mM led to a marked increase in the transition rate. Beyond 300 mM Naþ, however, further increase in Naþ concentration led to a reduction in rate. This nonmonotonic trend was observed for all Mg2þ concentrations examined, and indicates that the stability of the stacked structures is not a simple function of ionic strength. Evidently, at low concentrations, Na ions screen the interaction between Mg2þ and the junction, reducing the conformer

747

Conformational Species of Holliday Junction

The parallel conformation of the free DNA junction is not detectable

Figure 6. Competition and cooperation between monovalent and divalent ions. The dependence of transition rates for junction 7 on Naþ concentration in the presence of three different background Mg2þ concentrations of 2 (triangles), 5 (circles) and 20 mM (squares). Each rate is an average of ,five different single-junction molecules. Error bars reflect the heterogeneity in kCT values between molecules.

stability and hence accelerating the dynamics. At high concentrations, Na ions play a similar role as Mg2þ, and the two cooperate to stabilize the stacked structures.

Discussion The single-molecule studies confirm the dynamic nature of the four-way DNA junction, with constant exchange between stacked X-structures. They have also allowed us to examine the importance of species other than the majority antiparallel stacked conformers. While some of these exist in low populations, they are nevertheless of great potential significance in the functional role of the junction. The open intermediate of the nonmigratable junction Figure 3 shows a transition from open to stacked forms as Mg2þ concentration is increased. The fit shown is made to the Hill binding model with the Hill coefficient of 1. Although the fit cannot be interpreted as the site binding of one magnesium ion, the excellent agreement with data suggests that the transition from open to stacked species is relatively gradual, instead of being sharp and highly cooperative. Therefore, it is likely that the open state is significantly populated even in higher magnesium concentrations and is probably a stable intermediate during conformer transitions. However, we cannot estimate the lifetime of the open intermediate directly because of limited time resolution and signal level.

Original models of the Holliday junction were almost exclusively in a parallel conformation, and it therefore came as a surprise when the first structural data indicated that the four-way DNA junction in free solution preferentially adopted an antiparallel conformation.14,15 This bias was later shown to be true in the crystal too.20,22 Nevertheless, it is possible that in solution the junction might exist in a parallel conformation during some small fraction of the time, such that it would be undetectable in ensemble measurements. We therefore set out to try to detect this conformation by single-molecule spectroscopy. However, our single-molecule FRET data suggest that the fourway DNA junction does not exist in the parallel form, even for a short time. In the presence of 30 mM Mg2þ, the parallel form was not observable during a total time of 50 seconds with 5 ms time resolution. Therefore, a stable parallel conformation with a lifetime greater than 5 ms must be exceedingly rare (, 0.01%). It is still possible that parallel conformations exist with lifetimes shorter than 5 ms. With time resolution less than 5 ms, we did see short-lived high FRET states, which could correspond to short-lived parallel forms. But, as shown in Figure 4(c), (e) and (g), cross-correlation analysis with 1 –2 ms time resolution shows no anti-correlation. Also, at a low temperature (2 8C), the lifetimes of shortlived high FRET states of junction 1 did not increase (data not shown). From these observations, the short-lived high FRET states of less than 5 ms are unlikely to arise from a population of parallel forms, but are rather due to shot noise or some other artifact. The lack of parallel states does not appear to be a peculiarity of non-migratable junctions because our preliminary measurements on migratable junctions did not detect any evidence of parallel forms (S.A.M. et al., unpublished data). Thus, even if the parallel structure has a transient existence, it is unlikely to be important for the mechanism of genetic recombination unless stabilized by the interaction with proteins. Branch migration rate is likely to be limited by the rate of interconversion between conformers Branch migration is a key step in genetic recombination, involving the step-wise sequential exchange of base-pairing between the complementary strands of homologous duplexes. Two broad models have been suggested for the mechanism. Initially, the Holliday junction was believed to exist as a parallel structure, and it was proposed that simple rotational motion around the helical axes might induce branch migration without the need for unstacking the helices.40,41 However, it is now well established that four-way DNA junctions adopt an antiparallel conformation, and that

748

conclusion has now been reinforced and extendedhere. It therefore seems inescapable that branch migration must require an open conformation as an obligate intermediate, consistent with earlier observations28 that the rate is 1000-fold greater in the absence of Mg2þ. A recent atomic force microscopic study of surface-immobilized four-way DNA junctions in the absence of Mg2þ also showed that the open form must be achieved for branch migration to occur.42 Conformational interconversion between iso I and iso II conformers becomes slower as the concentration of Mg2þ increases. This is qualitatively similar to the behavior of spontaneous branch migration, the rate of which also decreases with increased Mg2þ concentration.36 Therefore, it is likely that the open structure is a common intermediate for branch migration and stacking conformer transitions.7 Because the frequency with which the open intermediate is accessed cannot be greater than the measured interconversion rate, this sets the upper limit on the stepping rate of branch migration, kBM. At high Mg2þ concentration, kCT is low, so the open intermediate is visited less frequently, and consequently kBM is reduced. For example, in the presence of 10 mM Mg2þ, kCT for junctions 1 and 7 is 47 s21 and 17 s21, respectively. Thus, the frequency with which the open inter21 mediate is visited, 4/(k21 I!II þ kII!I) sets the upper 21 21 limit as 30 s and 17 s on the kBM, with the caveat that this limit can be highly dependent on the local sequence. Here, the factor of 4 stems from the fact that the open state is likely to be visited four times on average during one cycle of conformer transition, twice during the actual transitions back and force and twice during a failed attempt that returns the molecule to the previous conformer. In the branch migration experiments of Panyutin & Hsieh28 kBM was measured as 3.9 s21 in the presence of 10 mM Mg2þ, which is much lower than the upper limits calculated from the data of junction 1 and 7 even though branch migration was measured at a higher temperature than conformer transitions (37 8C versus 20 8C). This indicates that during extended branch migration, multiple conformer transitions may occur before the branch point migrates a given step. Moreover, this does not appear to be a unique property of nonmigratable junctions (i.e. those incapable of branch migration), because our measurements on migratable junctions showed the expected behavior (S.A.M. et al., unpublished data). Local sequence and metal ions determine the conformer bias in the four-way DNA junction The conformational dynamics of four-way DNA junctions are affected by the base sequence at the point of strand exchange, and the nature and concentration of metal ions present (Table 1). Although junctions 1, 3 and 7 differ only in the identities of the nucleotides located immediately at the point of strand exchange, they exhibit markedly different

Conformational Species of Holliday Junction

conformational properties. In the presence of Mg2þ, junction 1 exists predominantly in the iso I stacking conformer, junction 3 favors iso II, while junction 7 favors neither significantly. The type of cation present may also affect the bias; while junction 7 has equal populations of iso I and iso II conformations in the presence of Mg2þ, iso I is favored in Naþ or hexammine cobalt (III) ions. However, the concentration of a given cation does not affect the conformer bias.

Conclusion Single-molecule studies underline the dynamic and polymorphic character of the four-way DNA junction. These species are in constant exchange between stacking conformers, where the position of equilibrium is determined first and foremost by the local sequence around the point of strand exchange together with some influence of cation concentration and type. Despite the dynamic character of these branched DNA species, our studies have shown that parallel forms of the DNA junction are not observed. Parallel forms with a lifetime greater than 5 ms are extremely rare (possibly nonexistent), with a probability lower than 0.01%. This contrasts with the corresponding RNA four-way junction, where a significant population of parallel forms exist.38 It is possible that the antiparallel form of the DNA junction is stabilized by favorable backbone –groove interactions,43,44 which are not possible in RNA due to the different geometry of the A-form helix. We could not directly detect the open form of the junction that must exist as an essential intermediate in the interconversion between stacking conformers. However, our data suggest that the open form is stably populated even in relatively high magnesium concentrations. The open form must be visited very frequently, hundreds of times a second in submillimolar magnesium concentrations, since stacking conformer transitions occur in a few milliseconds time scale in these physiologically relevant conditions. Therefore, the open form may play important roles in the mechanism of genetic recombination even though it is not the major species. This species must also be a common intermediate in the branch migration process. Relative rates indicate that conformer exchange equilibration will occur at each step of spontaneous branch migration. In summary, the four-way DNA junction intrinsically adopts an antiparallel stacked X-conformation that is in constant exchange between stacking conformers via a short-lived open intermediate. Any other behavior that might be required by the exigencies of recombination mechanisms must be conferred by protein binding. Indeed, it is striking that many proteins targeted to four-way DNA junctions (junction-resolving enzymes for example,45 – 49) do indeed distort the local and global structure.50 A clear example is

749

Conformational Species of Holliday Junction

provided in branch migration, accelerated by the RuvAB proteins in Escherichia coli.51 The RuvA protein stabilizes the structure of the junction in the unstacked square form,52 effectively driving the formation of a conformation with the properties of the open intermediate. Manipulation of DNA structure may therefore be one of the most important functions of recombination proteins, working to overcome the intrinsic conformational properties of the DNA junction.

Materials and Methods DNA preparation Vectors of junction 1 were prepared as following. Oligonucleotides of the following sequences were purchased from Integrated DNA Technologies. All the strands were purified by gel electrophoresis in polyacrylamide. J1b is 50 -Cy5-CCC TAG CAA GCC GCT GCT ACG G; J1Lb, 50 -CCC TAG CAA GCC GCT GCT ACG G; J1Rb, 50 -Cy3-CCC TAG CAA GCC GCT GCT ACG G; J1h, 50 -CCG TAG CAG CGA GAG CGG TGG G; J1Th, 50 -Biotin-CCG TAG CAG CGA GAG CGG TGG G; J1Qh, 50 -Cy5-CCG TAG CAG CGA GAG CGG TGG G; J1r, 50 -Biotin-CCC ACC GCT CTT CTC AAC TGG G; J1Tr, 50 -Cy5-CCC ACC GCT CTT CTC AAC TGG G; J1x, 50 Cy3-CCC AGT TGA GAG CTT GCT AGG G; J1Rx, 50 -Cy5-CCC AGT TGA GAG CTT GCT AGG G; J1Ax, 50 CCC AGT TGA GAG CTT GCT AGG G. Each junction is constructed by the hybridization of four strands: b, h, r and x. Junction 1-XB comprises strands J1b, J1h, J1r and J1x; junction 1-XR comprises J1Lb, J1Th, J1Tr and J1x; junction 1-BX comprises J1Rb, J1h, J1r and J1Rx; and junction 1-BR comprises J1Rb, J1Th, J1Tr and J1Ax. The four oligonucleotides were mixed with the ratio of 1 : 2 : 2 : 2 for Cy3-, Cy5-, biotin-labeled and unlabeled strands, respectively, in 10 mM Tris – HCl (pH 8.0), 50 mM NaCl, with a Cy3-strand concentration of approximately 2 mM. The mixture was cooled from 90 8C to 24 8C over three to four hours. Annealed junctions were stored at 220 8C. Junction 7 was prepared analogously.7

Data acquisition and analysis Data were acquired using software written in Visual Cþþ (Microsoft). They were analyzed using Origin (Origin Lab Corporation) and Matlab (The MathWorks, Inc.). The locations of junctions were manually identified in images scanned over 10 mm £ 10 mm area, by Cy5 signals. The fluorescence signals of the chosen molecules were recorded for 15 –30 seconds. Junctions with four perfectly base-paired arms should exhibit an anti-correlated behavior with respect to signals of Cy3 and Cy5 during conformational change or when Cy5 photoblinking/photobleached, and the photobleaching of Cy3 or Cy5 should be a one-step process. These criteria select 10 – 50% out of all data recorded. Rates of conformational fluctuations were obtained by a cross-correlation Ð analysis. The quantity, ID ðtÞIA ðt þ DtÞdt, was fitted by a single exponential function, expð2Dt=tÞ with the fitting parameter of correlation time, t. For each condition, about 5 – 20 molecules were analyzed. The data for each molecule were fitted to give the value of t. The set of t for a given condition gave the average and the standard deviation as represented in the Figures. The standard deviation of t reflects the statistical, local and possibly intrinsic heterogeneity of molecules in a given condition. Conformer transition rates, kCT were calculated from the correlation time using kCT ¼ t21 . The ratio between numbers of data points of each FRET state gives kI!II/kII!I. Values of the forward and backward transition rates (kI!II and kII!I) are obtained by the ratio of kI!II/kII!I in concert with the relation kCT ; kI!II þ kII!I.

Acknowledgements We thank A.-C. De´clais for providing junction 7; I. Rasnik, S. Myong, S. Hohng and R. Roy for advice and discussion on the experiment. C.J. thanks J. Milton for his guidance. Funding was provided by the NIH, NSF, Searle Scholars Award (T.H.) and by Cancer Research-UK (D.M.J.L.). S.A.M. was partially supported by the NIH molecular biophysics training grant and the NSF graduate fellowship.

Single-molecule measurements Details were as described.7 Briefly, junction molecules were immobilized to a streptavidin-coated glass surface and the donor fluorophores were excited using a solidstate 532 nm laser, with a power of 1 – 16 mW in a homebuilt confocal scanning microscope based on an inverted microscope with a 100 £ oil immersion objective with a numerical aperture of 1.4. A piezoelectric nanopositioner (Nano-Drive, Mad City Labs Inc.) was used to locate a single molecule under the laser focal spot, and fluorescence signals were detected by a pair of detectors (Photon counting module, SPCM-AQR-14, Perkin Elmer). Data were collected at 20(^2) 8C and the temperatures were monitored using a thermocouple. Unless otherwise specified, all measurements were made in a 10 mM Tris – HCl (pH 8.0), oxygen scavenger system (0.4% (w/v) glucose, 1% (v/v) 2-mercaptoethanol, 0.1 mg ml21 glucose oxidase 0.02 mg ml21 catalase) with specified amounts of MgCl2, NaCl, Co(NH3)6Cl3 and EDTA.

References 1. Cox, M. M., Goodman, M. F., Kreuzer, K. N., Sherratt, D. J., Sandler, S. J. & Marians, K. J. (2000). The importance of repairing stalled replication forks. Nature, 404, 37 – 41. 2. Holliday, R. (1964). A mechanism for gene conversion in fungi. Genet. Res. 5, 282– 304. 3. Lilley, D. M. J. (2000). Structure of helical junctions in nucleic acids. Quart. Rev. Biophys. 33, 109– 159. 4. Weiss, S. (2000). Measuring conformational dynamics of biomolecules by single molecule fluorescence spectroscopy. Nature Struct. Biol. 7, 724– 729. 5. Ha, T., Enderle, T., Ogletree, D. F., Chemla, D. S., Selvin, P. R. & 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.

750

6. Ha, T. (2001). Single molecule fluorescence resonance energy transfer. Methods, 25, 78. 7. McKinney, S. A., Declais, A. C., Lilley, D. M. J. & Ha, T. (2003). Structural dynamics of individual Holliday junctions. Nature Struct. Biol. 10, 93 – 97. 8. Ying, L., Green, J. J., Li, H., Klenerman, D. & Balasubramanian, S. (2003). Studies on the structure and dynamics of the human telomeric G quadruplex by single-molecule fluorescence resonance energy transfer. Proc. Natl Acad. Sci. USA, 100, 14629– 14634. 9. Rothwell, P. J., Berger, S., Kensch, O., Felekyan, S., Antonik, M., Wohrl, B. M. et al. (2003). Multiparameter single-molecule fluorescence spectroscopy reveals heterogeneity of HIV-1 reverse transcriptase: primer/template complexes. Proc. Natl Acad. Sci. USA, 100, 1655– 1660. 10. Rasnik, I., Myong, S., Cheng, W., Lohman, T. M. & Ha, T. (2004). DNA-binding orientation and domain conformation of the E. coli Rep helicase monomer bound to a partial duplex junction: single-molecule studies of fluorescently labeled enzymes. J. Mol. Biol. 336, 395– 498. 11. Ha, T., Rasnik, I., Cheng, W., Babcock, H. P., Gauss, G., Lohman, T. M. & Chu, S. (2002). Initiation and reinitiation of DNA unwinding by the Escherichia coli Rep helicase. Nature, 419, 638– 641. 12. Katiliene, Z., Katilius, E. & Woodbury, N. W. (2003). Single molecule detection of DNA looping by NgoMIV restriction endonuclease. Biophys. J. 84, 4053– 4061. 13. Clegg, R. M., Murchie, A. I. H. & Lilley, D. M. J. (1994). The solution structure of the four-way DNA junction at low-salt conditions—a fluorescence resonance energy transfer analysis. Biophys. J. 66, 99 – 109. 14. Duckett, D. R., Murchie, A. I., Diekmann, S., von Kitzing, E., Kemper, B. & Lilley, D. M. (1988). The structure of the Holliday junction and its resolution. Cell, 55, 79 – 89. 15. Murchie, A. I., Clegg, R. M., von Kitzing, E., Duckett, D. R., Diekmann, S. & Lilley, D. M. (1989). Fluorescence energy transfer shows that the four-way DNA junction is a right-handed cross of antiparallel molecules. Nature, 341, 763– 766. 16. Cooper, J. P. & Hagerman, P. J. (1987). Gel electrophoretic analysis of the geometry of a DNA fourway junction. J. Mol. Biol. 198, 711 – 719. 17. Gough, G. W. & Lilley, D. M. (1985). DNA bending induced by cruciform formation. Nature, 313, 154– 156. 18. Clegg, R. M., Murchie, A. I., Zechel, A., Carlberg, C., Diekmann, S. & Lilley, D. M. (1992). Fluorescence resonance energy transfer analysis of the structure of the four-way DNA junction. Biochemistry, 31, 4846– 4856. 19. Churchill, M. E., Tullius, T. D., Kallenbach, N. R. & Seeman, N. C. (1988). A Holliday recombination intermediate is twofold symmetric. Proc. Natl Acad. Sci. USA, 85, 4653– 4656. 20. Eichman, B. F., Vargason, J. M., Mooers, B. H. M. & Ho, P. S. (2000). The Holliday junction in an inverted repeat DNA sequence: sequence effects on the structure of four-way junctions. Proc. Natl Acad. Sci. USA, 97, 3971– 3976. 21. Nowakowski, J., Shim, P. J., Prasad, G. S., Stout, C. D. & Joyce, G. F. (1999). Crystal structure of an 82-nucleotide RNA– DNA complex formed by the 10 – 23 DNA enzyme. Nature Struct. Biol. 6, 151– 156. 22. Ortiz-Lombardia, M., Gonzalez, A., Eritja, R.,

Conformational Species of Holliday Junction

23. 24. 25.

26.

27.

28. 29.

30.

31. 32.

33.

34.

35.

36. 37. 38. 39.

40.

Aymami, J., Azorin, F. & Coll, M. (1999). Crystal structure of a DNA Holliday junction. Nature Struct. Biol. 6, 913– 917. Grainger, R. J., Murchie, A. I. H. & Lilley, D. M. J. (1998). Exchange between stacking conformers in a four-way DNA junction. Biochemistry, 37, 23 – 32. Meselson, M. S. & Radding, C. M. (1975). A general model for genetic recombination. Proc. Natl Acad. Sci. USA, 72, 358– 361. Miick, S. M., Fee, R. S., Millar, D. P. & Chazin, W. J. (1997). Crossover isomer bias is the primary sequence-dependent property of immobilized Holliday junctions. Proc. Natl Acad. Sci. USA, 94, 9080– 9084. Murchie, A. I., Portugal, J. & Lilley, D. M. (1991). Cleavage of a four-way DNA junction by a restriction enzyme spanning the point of strand exchange. EMBO J. 10, 731– 738. Overmars, F. J. J. & Altona, C. (1997). NMR study of the exchange rate between two stacked conformers of a model Holliday junction. J. Mol. Biol. 273, 519– 524. Panyutin, I. G. & Hsieh, P. (1994). The kinetics of spontaneous DNA branch migration. Proc. Natl Acad. Sci. USA, 91, 2021– 2025. Thompson, B. J., Camien, M. N. & Warner, R. C. (1976). Kinetics of branch migration in doublestranded DNA. Proc. Natl Acad. Sci. USA, 73, 2299– 2303. Orr-Weaver, T. L., Szostak, J. W. & Rothstein, R. J. (1981). Yeast transformation: a model system for the study of recombination. Proc. Natl Acad. Sci. USA, 78, 6354– 6358. Chen, Y. J., Yu, X. & Egelman, E. H. (2002). The hexameric ring structure of the Escherichia coli RuvB branch migration protein. J. Mol. Biol. 319, 587– 591. Constantinou, A., Davies, A. A. & West, S. C. (2001). Branch migration and Holliday junction resolution catalyzed by activities from mammalian cells. Cell, 104, 259– 268. Iwasaki, H., Takahagi, M., Nakata, A. & Shinagawa, H. (1992). Escherichia coli RuvA and RuvB proteins specifically interact with Holliday junctions and promote branch migration. Genes Dev. 6, 2214– 2220. Muller, B., Tsaneva, I. R. & West, S. C. (1993). Branch migration of Holliday junctions promoted by the Escherichia coli RuvA and RuvB proteins. I. Comparison of RuvAB- and RuvB-mediated reactions. J. Biol. Chem. 268, 17179– 17184. van Gool, A. J., Shah, R., Mezard, C. & West, S. C. (1998). Functional interactions between the holliday junction resolvase and the branch migration motor of Escherichia coli. EMBO J. 17, 1838– 1845. Panyutin, I. G., Biswas, I. & Hsieh, P. (1995). A pivotal role for the structure of the Holliday junction in DNA branch migration. EMBO J. 14, 1819– 1826. Duckett, D. R., Murchie, A. I. & Lilley, D. M. (1990). The role of metal ions in the conformation of the four-way DNA junction. EMBO J. 9, 583– 590. Hohng, S., Wilson, T. J., Tan, E., Clegg, R. M., Lilley, D. M. & Ha, T. (2004). Conformational flexibility of four-way junctions in RNA. J. Mol. Biol. 336, 69 – 79. Gessner, R. V., Quigley, G. J., Wang, A. H., van der Marel, G. A., van Boom, J. H. & Rich, A. (1985). Structural basis for stabilization of Z-DNA by cobalt hexaammine and magnesium cations. Biochemistry, 24, 237– 240. Sigal, N. & Alberts, B. (1972). Genetic recombination: the nature of a crossed strand-exchange between two

751

Conformational Species of Holliday Junction

41. 42.

43.

44. 45. 46.

homologous DNA molecules. J. Mol. Biol. 71, 789–793. Meselson, M. (1972). Formation of hybrid DNA by rotary diffusion during genetic recombination. J. Mol. Biol. 71, 795–798. Lushnikov, A. Y., Bogdanov, A. & Lyubchenko, Y. L. (2003). DNA recombination: holliday junctions dynamics and branch migration. J. Biol. Chem. 278, 43130– 43134. von Kitzing, E., Lilley, D. M. J. & Diekmann, S. (1990). The stereochemistry of a four-way DNA junction: a theoretical study. Nucl. Acids Res. 18, 2671–2683. Timsit, Y. & Moras, D. (1991). Groove-backbone interaction in B-DNA. Implication for DNA condensation and recombination. J. Mol. Biol. 221, 919– 940. Bennett, R. J. & West, S. C. (1995). Structural analysis of the RuvC-Holliday junction complex reveals an unfolded junction. J. Mol. Biol. 252, 213– 226. Duckett, D. R., Panis, M. & Lilley, D. M. J. (1995). Binding of the junction-resolving enzyme bacteriophage T7 endonuclease I to DNA—separation of

47.

48. 49. 50. 51. 52.

binding and catalysis by mutation. J. Mol. Biol. 246, 95 – 107. Po¨hler, J. R. G., Giraud-Panis, M.-J. E. & Lilley, D. M. J. (1996). T4 endonuclease VII selects and alters the structure of the four-way DNA junction; binding of a resolution-defective mutant enzyme. J. Mol. Biol. 260, 678– 696. White, M. F. & Lilley, D. M. J. (1997). The resolving enzyme CCE1 of yeast opens the structure of the four-way DNA junction. J. Mol. Biol. 266, 122– 134. De´clais, A.-C. & Lilley, D. M. J. (2000). Extensive central disruption of a four-way junction on binding CCE1 resolving enzyme. J. Mol. Biol. 296, 421– 433. Lilley, D. M. J. & White, M. F. (2001). The junctionresolving enzymes [Review]. Nature Rev. Mol. Cell Biol. 2, 433– 443. West, S. C. (1997). Processing of recombination intermediates by the RuvABC proteins [Review]. Annu. Rev. Genet. 31, 213– 244. Parsons, C. A., Stasiak, A., Bennett, R. J. & West, S. C. (1995). Structure of a multisubunit complex that promotes DNA branch migration. Nature, 374, 375– 378.

Edited by P. J. Hagerman (Received 24 February 2004; received in revised form 24 May 2004; accepted 7 June 2004)

Exploring Rare Conformational Species and Ionic ...

how different enzymes recognize and process the .... data points corresponding to each FRET state (see .... In the branch migration process the open structure.

735KB Sizes 0 Downloads 162 Views

Recommend Documents

MN River Valley Master Plan Rare Species--January ...
Shorelines of lakes, ponds, and slow moving streams. Water-hyssop. Bacopa rotundifolia. Aquatic Plant. THR. Small rainwater pools on bedrock outcrops in western Minnesota. American Bittern. Botaurus lentiginosus. Bird ... Mixed unit or sequence (arch

Synthesis, conformational studies and inclusion ...
spectra were recorded on a Voyager-DETM STR Bio-. spectrometry_Workstation made by Applied Biosys- tems Inc. Azacalixarenes 1 [7d, l] and 5 [7b] and o,o'- bis(hydroxymethyl) p-substituted phenol dimers 9 [7d, l] and 10 [7b] were prepared according to

1 STRUCTURE AND CONFORMATIONAL CHANGES ...
domain, we have developed a FRET approach .... The 100 µL fluorophore-labeled receptors were then separated .... 100 µM TCEP and 3 equivalents of free dye.

Estimating fishing mortality of major target species and species ... - frdc
Background. The volume of shark ..... channels (newspapers, fishing websites and newsletters) and word-of-mouth. Incentives including ...... images these should be outlined in this section outline and attach them where possible. Manuscript ...

Estimating fishing mortality of major target species and species ... - frdc
improve data quality; and the designation of a number species as either no-take or .... management of shark mortality needs to consider all interactions.

Conformational diversity and protein evolution – a 60 ...
binding kinetics provided the first data indicating pre- existing isomers in .... duplication and mutation could then provide the increased genetic diversity to drive ...

Conformational Proofreading: The Impact of ...
May 23, 2007 - Kij is the dissociation constant of the complex formed from the i-th ligand conformation ... depends on the concentrations of the complexes, which depend on Kij, and on the functionality of each ..... Johnson KA (1993) Conformational c

Spontaneous conformational change and toxin binding ...
Jun 17, 2008 - E-mail: [email protected]. This article contains ..... Schwede T, Kopp J, Guex N, Peitsch MC (2003) SWISS-MODEL: An automated protein.

Rare GK.pdf
iao1. mawwng. LSLl6)1C8 OJ6)0 3 oi(O36115 Q'D amo 6)6). 6flJO6G)1 0Efl arJ( (J16)006)6)?. - 1300 a61)J00)o, 0CT5QJ0(T. 6)6)flJ-6)6)riJJ ri110 ()JW1 O0)0.

Mapping the Native Conformational Ensemble of Proteins from a ...
May 23, 2013 - combination with an accurate atomic-level computational modeling, can disclose the details of protein behavior. We here propose a fast and efficient protocol employing molecular dynamics (MD) simulations and NMR chemical shifts, which

rare disease asean.pdf
Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. rare disease asean.pdf. rare disease asean.pdf. Open. Extract.

Reactive oxygen species, Aging and Antioxidative Nutraceuticals.pdf ...
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. Reactive oxygen ...

A circular dichroism–DFT method for conformational study of ... - Arkivoc
Dec 27, 2016 - either acyclic (1–5) or cyclic (6–10) and the ester groups are separated either by two, three, or four carbon atoms. Note that ..... Rigid, three- and four-membered-ring derivatives as well as the more flexible ...... Montgomery, J

Ionic Bonds and Formulas Excerpt.pdf
given the symbol of a main group element, indicate the most likely number of electrons the atom will. gain or lose. • predict the charge on ions from the electron affinity, ionization energies, and electron configuration. of the atom. • describe

Why Coincidences, Miracles, and Rare Events Happen ...
Coincidences, Miracles, and Rare Events Happen ... and universe we live in, "The Improbability Principle" will transform how you think about serendipity and ...

Rare and Uncommon Animals of Vermont, Nov. 2013.pdf
Wildlife Department, 1 National Life Drive North 2, Montpelier, VT 05620-3702. .... 2013.pdf. Vermont - Rare and Uncommon Animals of Vermont, Nov. 2013.pdf.

Raine Miller, Blackstone 4, Rare And Precious Things.pdf
Page 3 of 289. Raine Miller, Blackstone 4, Rare And Precious Things.pdf. Raine Miller, Blackstone 4, Rare And Precious Things.pdf. Open. Extract. Open with.

Reactive oxygen species, Aging and Antioxidative Nutraceuticals.pdf
Reactive oxygen species, Aging and Antioxidative Nutraceuticals.pdf. Reactive oxygen species, Aging and Antioxidative Nutraceuticals.pdf. Open. Extract.