Anda di halaman 1dari 8

Methods 47 (2009) 206–213

Contents lists available at ScienceDirect

Methods
journal homepage: www.elsevier.com/locate/ymeth

AFM for analysis of structure and dynamics of DNA and protein–DNA complexes
Yuri L. Lyubchenko *, Luda S. Shlyakhtenko
Department of Pharmaceutical Sciences, College of Pharmacy, COP 1012, University of Nebraska Medical Center, 986025 Nebraska Medical Center, Omaha, NE 68198-6025, USA

a r t i c l e i n f o a b s t r a c t

Article history: This paper describes protocols for studies of structure and dynamics of DNA and protein–DNA complexes
Accepted 10 September 2008 with atomic force microscopy (AFM) utilizing the surface chemistry approach. The necessary specifics for
Available online 7 October 2008 the preparation of functionalized surfaces and AFM probes with the use of silanes and silatranes, includ-
ing the protocols for synthesis of silatranes are provided. The methodology of studies of local and global
Keywords: conformations DNA with the major focus on the time-lapse imaging of DNA in aqueous solutions is illus-
Atomic force microscopy (AFM) trated by the study of dynamics of Holliday junctions including branch migration. The analysis of nucle-
Force microscopy
osome dynamics is selected as an example to illustrate the application of the time-lapse AFM to studies of
DNA dynamics
Local DNA structures
dynamics of protein–DNA complexes. The force spectroscopy is the modality of AFM with a great impor-
Holliday junctions tance to various fields of biomedical studies. The AFM force spectroscopy approach for studies of specific
DNA–protein interactions protein–DNA complexes is illustrated by the data on analysis of dynamics of synaptic SfiI–DNA com-
Protein–protein interactions plexes. When necessary, additional specifics are added to the corresponding example.
Single-molecule techniques Ó 2008 Elsevier Inc. All rights reserved.
Solution time-lapse imaging

1. Introduction surfaces with desirable characteristics. The latter property was crit-
ical for routine and reliable immobilization of the sample for AFM
Single molecule biophysics technologies due to their capability force spectroscopy [23]. This paper outlines specifics for the prepa-
to detect transient states of molecules and biomolecular complexes ration of surfaces, samples for AFM studies including the time-lapse
are the methods of choice for studies DNA structure and dynamics. AFM and AFM force spectroscopy. A few examples illustrating the
Atomic force microscopy (AFM) is one of these methods. A unique advantages of the described techniques for the study of DNA struc-
feature of this instrument is its capability to operate in two different ture and dynamics analysis are provided. However, this is a short
modalities, topographic imaging and measuring of intermolecular list. In fact, the APS–mica methodology has been used in the AFM
interactions. AFM was also successfully applied to studies of nucleic topographic studies of alternative DNA structures such as cruci-
acids. Part of this success is due to the development of reliable sam- forms, three-way junctions, open DNA regions (see reviews
ple preparation procedures. Several such methods were developed [20,24] and references therein) and various protein–DNA com-
simultaneously in a number of laboratories [1–10]. A major feature plexes [25–30]. In addition, the silatrane chemistry including
of these methods is the use of a specially prepared surface that holds APS–mica methodology was used in numerous topographic and
(usually electrostatically) the sample in place during scanning. We the force spectroscopy studies in the area of protein misfolding
used aminosilanes [10–15] and silatranes [16–22] to functionalize and aggregation [22,31–39].
the mica surface with amine groups. The major advantage of these
sample preparation procedures is that they work under a wide vari- 2. Experimental design
ety of ionic conditions, pH and over a wide range of temperatures.
These characteristics of functionalized mica allowed us to image nu- 2.1. Materials
cleic acids (DNA, dsRNA, kDNA) and nucleoprotein complexes of dif-
ferent types (reviewed in [10,13,20]). Once prepared, samples are 2.1.1. Chemicals
stable and do not absorb any contaminants for weeks with minimal Commercially available 3-aminopropyltriethoxy silane (e.g., Flu-
precautions for storing. An important property of silatrane, such as ka, Chemika-BioChemika, Switzerland, Aldrich, USA, United Chem-
the possibility to synthesize functionalized silatranes capable of ical Technology, USA) and N,N-diisopropylethylamine (Aldrich,
bringing to the surface a target group including chemically reactive Sigma). It is recommended to redistill APTES and store under argon.
group, is a unique property of silatranes, allowing to prepare the
2.1.2. Mica substrate
* Corresponding author. Fax: +1 402 559 9543. Any type of commercially available mica sheets (green or ruby
E-mail address: ylyubchenko@unmc.edu (Y.L. Lyubchenko). mica). Asheville-Schoonmaker Mica Co (Newport News, VA)

1046-2023/$ - see front matter Ó 2008 Elsevier Inc. All rights reserved.
doi:10.1016/j.ymeth.2008.09.002
Y.L. Lyubchenko, L.S. Shlyakhtenko / Methods 47 (2009) 206–213 207

supplies with thick and large (more than 5  7 cm) sheets suitable OH
for making the substrates of different sizes. Mica APS
O OH
2.1.3. Water OH OH O Water Si N
Si O Si O Si H2N O O
Double glass distilled or deionized water filtered through O
0.5 lm filter. Si O Si O Si

2.2. Functionalization of mica with 3-aminopropyltriethoxy silane Fig. 1. Scheme for reaction of 3-aminopropylsilatrane (APS) with hydroxyl groups
on a mica (silicon) surface. The initial adduct reacted with one hydroxyl group can
(AP–mica preparation)
reach with a second surface OH group forming the indicated product in a reversible
equilibrium.
The protocol below describes the methodology for the prepara-
tion of positively charged mica surface provided by covalent bind-
ing of APTES to mica surface. ator. The shelf life of the stock solution is not less than 6 months.
Place two plastic caps (cut them from regular 1.5 ml plastic Prepare working APS solution for mica modification dissolving
tubes) on the bottom of 2 l desiccators and vacuum it with a regu- the stock solution in 1:300 ratio in water; it can be stored at room
lar vacuum pump and fill with argon. Cleave mica sheets (approx. temperature for several days. Cleave mica sheets of needed sizes
5  5 cm) to make them as thin as 0.1–0.05 mm and mount at the (typically 1  3 cm) to make them as thin as 0.05–0.1 mm, place
top of the desiccators. Place 30 ll of APTES into one plastic cap and them in appropriate plastic tubes and pour working APS solution
10 ll of N,N-diisopropylethylamine (Aldrich) into another cap and to cover the mica sheet and leave on the bench for 30 min. Remove
allow the reaction to proceed for 1–2 h. After that remove the cap the mica sheets wash with deionized water and dry them under
with APTES and purge argon for 2 min. Leave the sheets for 1–2 Argon stream. The strips are ready for the sample preparation.
days in the desiccator to cure; AP–mica is ready for the sample Note, however, as prepared, the APS–mica sheets can be stored un-
deposition after that. This procedure allows one to obtain a weak der Argon for several days.
cationic surface with rather uniform distribution of the charge.
Dry argon atmosphere is crucial for obtaining the substrates for 2.5. Sample preparation for AFM imaging in air
AFM studies and for storage of the substrate. Allow the gas to flow
while you open the desiccator. With these precautions the AP– Two procedures are described. AP–mica is mentioned in the
mica substrates retain their activity for several weeks. protocols below, in the droplet and immersion procedures, how-
ever, the same protocol is also applied for APS–mica. The first pro-
2.3. Synthesis of 1-(3-aminopropyl)silatrane (APS) cedure is attractive due to the use of small amount of the sample,
whereas the immersion procedure is recommended for procedures
APS is not commercially available, but can be synthesized with requiring the control of environmental conditions such as temper-
use of rather simple equipment. A catalytic amount of sodium me- ature, pH and salt concentration.
tal (5 mg) is added to 15.0 ml (16.8 g, 0.11 mol) of triethanolamine
(Aldrich) in a 250 ml round-bottom flask under argon or nitrogen 2.5.1. The droplet procedure
atmosphere and allowed to form a solution. A rubber balloon is Prepare the solution of the sample (DNA, RNA, protein–DNA
then attached to the flask via a rubber septum and a needle to al- complex) in appropriate buffer. DNA concentration should be be-
low hydrogen to escape without building up pressure. Moderate tween 0.1 and 0.01 lg/ml depending on the size of the molecules.
heat (up to 100 °C) can be applied to accelerate the process, but Place 5–10 ll of the solution in the middle of AP–mica substrate
the mixture should be allowed to cool to room temperature before (usually 1  1 cm2) for 2–3 min. Rinse the surface thoroughly with
the next step. An equivalent amount of (3-aminopropyl)triethox- water (2–3 ml per sample) to remove all buffer components. Five
ysilane (26.4 ml or 25.0 g, 0.11 mol) is added to the mixture, then millilitre plastic syringe is very useful for rinsing. Attach an appro-
the flask is placed into a 60 °C water bath and connected to the vac- priate plastic tip instead of metal needle. Dry the sample by blow-
uum line to absorb ethanol released in the reaction. This reaction ing with clean argon gas. The sample is ready for imaging. Store the
can be simply performed on a rotary evaporator. At the end of samples in vacuum cabinets or desiccators filled with argon.
the reaction, the mixture loses approximately 17 g of ethanol and
is turned into a solid. This process can take more than 24 h, but 2.5.2. The immersion procedure
mechanical or occasional manual stirring can reduce time to This procedure is recommended if the deposition should be per-
1–2 h. Evaporation at the end with 150 ml of xylenes at 60 °C can formed at strictly controlled temperature conditions (0 °C or ele-
also accelerate the process and help crystallization. Vacuum-dried vated temperatures) and long-time incubations. Preincubate the
APS obtained by this method can be used directly for AFM, or for DNA solution for 10–20 min to allow the temperature to equili-
better results and higher stability, the product can be purified by brate. Immerse a piece of AP–mica into the vials and leave it for
crystallization. Minute amounts of sodium hydroxide in the prod- 10–20 min to allow the samples to adsorb to the surface. Remove
uct does not practically affect the performance of the reagent, or the specimen, rinse with water thoroughly and dry under the ar-
change the pH of stock solutions of APS used for AFM. Recrystalli- gon flow. The sample is ready for imaging. The samples can be
zation from xylenes provides 20 g (80% yield) of colorless solid stored in vacuum cabinet or under argon.
powder. The synthesized APS has the following characteristics:
m.p. 91–94 °C (open capillary tube); m.p. 87.2–87.9 °C (sealed cap- 2.6. Procedure for AFM imaging in air
illary tube). 1H NMR (DMSO-d6), ppm: 0.08–0.14 (2H, m, SiCH2);
1.1 (2H, br. s, NH2); 1.28–1.37 (2H, m, CH2); 2.37 (2H, t J = 7.2 Hz, The procedure for imaging in air is straightforward: mount the
NCH2); 2.77 (6H, t J = 5.9 Hz, NCH2); 3.59 (6H, t J = 5.9 Hz, OCH2). sample and start approaching. Although both contact and intermit-
tent (tapping) modes can be used, the latter is preferable and al-
2.4. Mica functionalization with APS lows one to get images of DNA and DNA–protein complexes
routinely. Any types of probes designed for non-contact imaging
Fig. 1 schematically illustrates the reaction of APS with mica. can be used. NanoProbe TESP tips (Digital Instruments, Inc.) or
Prepare 50 mM APS stock solution in water and store it in refriger- Olympus and conical sharp silicon probes of K-TEK International
208 Y.L. Lyubchenko, L.S. Shlyakhtenko / Methods 47 (2009) 206–213

(Portland, OR) work well. Typical tapping frequency of 240– The immobilization chemistry is shown in Fig. 2. A freshly
380 kHz; scanning rate of 1.5–2.5 Hz allows one to obtain stable cleaved mica surface (cartoon 1) is treated with an aqueous solu-
images. tion of maleimide silatrane (MAS, 2) for several minutes and then
rinsed, leading to a surface containing immobilized maleimide
2.7. Procedures for imaging in solution units as depicted in cartoon 3. Surface 3 is then treated with a buf-
fered solution containing the solution of thiol-modified DNA. The
The capability of AFM to perform scanning in liquid is its most terminal thiol group covalently adds to the maleimide units on
attractive feature for numerous biological applications allowing the surface leading to the immobilized DNA as shown in cartoon 4.
imaging at conditions close to physiological ones. In addition, this Thiol group was incorporated into DNA during a standard
mode of imaging permits one to eliminate undesirable resolution- chemical synthesis. Specifics for the preparation and purification
limiting capillary effect typical for imaging in air (e.g., [8,40,41]). As of DNA can be found in [45]. To reduce protected thiol groups of
a result, images of DNA filaments as thin as 3 nm were obtained the oligonucleotides, the duplexes were treated with 10 mM buf-
in water solutions [42] and helical periodicity was observed when fered solution (10 mM HEPES, 50 mM NaCl, pH 7.0) of TCEP-hydro-
dried DNA samples were imaged in propanol [11]. In addition to chloride (Tris(2-carboxyethyl)phosphine) at 25 °C for 10 min.
APS–mica, our previously developed procedure with the use of Silicon nitride (Si3N4) or MLCT AFM tips were initially washed in
AP–mica can be used as a substrate for imaging in liquid [10]; note ethanol by immersion for 30 min and then activated by UV treat-
that the first images of DNA in fully hydrated state were obtained ment for 30 min t. Activated tips and freshly cleaved mica surface
by the use of AP–mica [11]. This type of imaging is recommended were treated with 167 lM solution of maleimide silatrane for 3 h
in cases when dynamics is studied. The following procedure is de- followed by rinsing with deionized water. Maleimide functional-
scribed for the use of MultiMode AFM (Veeco, Santa Barbara, CA) ized AFM tips and mica were incubated for 1 h, respectively, in
and can be easily adapted for other systems. Install an appropriate 10 lM and 80 lM of double stranded DNA. After washing with
tip in the holder designed for imaging in liquid (fluid cell). Use HEPES buffer solution (10 mM HEPES, 50 mM NaCl, pH 7.0), unre-
Si3N4 big triangle with thick legs or small triangle with thin legs acted maleimide moieties were quenched with buffered 10 mM
cantilevers [42]. Mount the AP–mica or APS–mica sheet on the b-mercaptoethanol solution by 10 min treatment at room temper-
stage of the microscope. Mica pieces of 1  1 cm are sufficient for ature. The modified tips and mica were washed with 10 mM
MultiMode AFM design of fluid cell. Attach the head of the micro- HEPES, 50 mM NaCl, pH 7.0 and stored in the same buffer until use.
scope with installed fluid cell and make appropriate adjustments of
the microscope. Approach the sample to the tip manually, leaving 3. Specific AFM experiments with the use of functionalized
50–100 lm gap between the tip and the surface. This gap is suf- surfaces
ficient to inject 50 ll of the sample. Allow the microscope to ap-
proach the sample and to engage the tip. Adjust the values for the 3.1. AFM imaging with the use of functionalized surfaces
set point voltage and drive amplitude to improve the quality of
images and start the data acquisition using the continuous mode The procedure for the preparation of samples with the use of
scanning. functionalized AP–mica or APS–mica procedures is straightfor-
The best resonance frequencies are usually around 6–9 kHz. APS ward. The sample is placed onto the surface to allow the sample
modified mica also can be used in MFP 3D instrument (Asylum Re- to diffuse to the surface and bind. After that, the sample is rinsed
search, Santa Barbara, CA). For that, the piece of mica should be with water, dried with a clean gas and placed on the microscope
glued to the glass slide first and then modified as described above stage for imaging. Fig. 3 shows typical images obtained with the
with APS solution. use of described protocols. We were able to image as well such
an alternative DNA structure as supercoil stabilized cruciforms
2.8. AFM force spectroscopy [46–48]. The following features of AP– and APS–mica protocol
were critical in reliable and reproducible detection of these
This section outlines specifics for the immobilization of mole- dynamic structures.
cules utilizing maleimide silatrane (MAS) methodology [43], the
suitability of which has been proven in a number of single mole- 1. The samples can be deposited in a wide range of ionic condi-
cule experiments including AFM force spectroscopy [22,43–45]. tions, temperatures and pH.

OH
2 (Maleimide silatrane, MAS)
O O OH
O
O Si N
N N O O
OH OH OH 4 OH
Si O Si O Si O O Si O Si O Si
water
1 (mica surface cartoon) 3 (mica surface covalently modified with maleimide-PEG-silatrane units)

Thiolate DNA in buffer OH

O O OH
O
O Si N
S N N O O
4 OH
DNA
O O Si O Si O Si

4 (cartoon showing immobilized amyloid peptide on mica by way of a C-terminal cys)

Fig. 2. Immobilization of thiolated DNA on the functionalized MAS–mica surface.


Y.L. Lyubchenko, L.S. Shlyakhtenko / Methods 47 (2009) 206–213 209

Fig. 3. AFM images of plasmid DNA with two types of alternative structures stabilized by negative supercoiling. (A) Bulges induced by denaturation at (ATTCT)29- region [1].
Open regions are indicated with arrows. (B) H-DNA appearing as a thick stem indicated with an arrow [26]. The samples were obtained by deposition on APS–mica.

2. No divalent cations are needed for the sample binding to the plexes and specifics for these experiments can be found in a num-
surface. ber of papers [25–28,59,30].
3. The substrate is stable and retains the binding activity during
several days after the preparation. 3.2. AFM imaging in aqueous solutions
4. The samples deposited on the substrates are stable and do not
accumulate impurities during several months with minimal Time-lapse imaging mode is one of the attractive modes of
precautions for the sample storage. topographic AFM applications. In the pioneering work [60], AFM
was applied to the imaging the fibrin self assembly. In the series
The following issue related to the effect of the surface on the of works from the group of Bustamante, AFM was applied to the
DNA conformation should be taken into a consideration. Deposi- observation of RNA polymerase movement along DNA [8,9,61–
tion of DNA molecules on a surface certainly leads to distortion 63]. We used time-lapse imaging and AP–mica methodology for
of its structure, but the extent of such distortion depends on the studies dynamics of local and global dynamics of supercoiled
physical and chemical characteristics of the surface and the DNA with the focus on structural dynamics of cruciforms [42,47].
DNA–surface interactions. Quite popular methods utilizing the However, for the experiments in aqueous solutions APS–mica,
use of divalent cations work very well for imaging of linear DNA there appears to be a better substrate than AP–mica. We observed
[1,49,50]. At the same time, the use of a modified Mg-assisted tech- that the APS–mica surface remains smooth during experiments in
nique the images of supercoiled DNA appear with large loops be- liquid, rendering APS–mica superior to AP–mica [18]. Controlling
tween the very tightly twisted segments of the plectonemic the surface density of amines is another important advantage of
superhelix was a typical feature of such images [51–55]. Such large the APS–mica method for the time-lapse AFM imaging experi-
loops should not present in DNA in solution and they are not ap- ments. These two features were critical, allowing single molecule
pear on the computer simulated images [56]; therefore the appear- AFM detection of the extensive dynamics of three-way DNA junc-
ances of large loops in supercoiled DNA are probably induced by tions [16], observations consistent with our earlier DNA cyclization
the immobilization procedure. The use of functionalized mica al- studies in solution [64,65]. The observation of the H-to-B-form
lowed us to eliminate or at least to minimize above-mentioned transition [18,24] is another example. The results on direct visual-
problems and to obtain the images of supercoiled DNA [42] fully ization of recombination DNA dynamics enabling us to test the
consistent with the data obtained in solution [57] and the predic- models for branch migration [20,66] along with visualization of
tions of theoretical analysis for the conformations of supercoiled dynamics of nucleosomes are briefly described in subsections
DNA [56]. In addition, supercoiled DNA in low salt buffer, but in below.
the presence of Mg2+ cations are observed as tightly interwound
plectonemic molecules [19,42]. This observation is in perfect 3.2.1. Dynamics of Holliday junctions and branch migration
agreement with data in solution [58], suggesting that the charge In time-lapse AFM experiments, the sample was injected into an
neutralization effect of Mg2+ cations is much stronger than that AFM flow cell mounted on APS–mica and imaged with an AFM con-
of monovalent cations. tinuously without drying the sample. A selected area with unam-
Altogether, these data further warrant the application of the biguously identified Holliday junctions (HJs) was scanned
functionalized mica techniques to studies of global DNA conforma- continuously and the buffer change was performed without inter-
tions and the structure and dynamics of negative DNA supercoil- ruption of the scanning. The results for one such a time-lapse
ing, which is a natural state of DNA at physiological conditions experiment are shown in Fig. 4. The arms subsequently move to
and within cells in particular. Local DNA structures are very labile an extended conformation (frame 9) followed by branch migration
and dynamic, so such conditions as ionic strength and pH are crit- [12,13], resulting in the dissociation of the molecule into two linear
ical for their visualization. Furthermore, our AFM study revealed strands as imaged in frame 13. A complete data set illustrated
the interplay between the conformational transition of a cruciform dynamics of the HJ and branch migration for this and other mole-
and the global conformation of supercoiled DNA can be a molecu- cules shown as movies can be viewed in the supplement to paper
lar trigger for the slithering of DNA chains, the process that facili- [66]. Quantitative analysis of the time-lapse data led us to con-
tates communication between distant sites in the molecule clude that branch migration requires unfolding of HJ, whereas fold-
[24,48]. Note that both AP– and APS–mica have very similar char- ing of the junction into conformations with antiparallel or parallel
acteristics in terms of the surface smoothness, holding the sample orientation of exchanging arms terminates branch migration.
on the surface provided and the high reproducibility of the results. Time-lapse AFM with the use of APS–mica was recently applied
Both surfaces were very useful for studies of protein–DNA com- to the analysis of the role of DNA supercoiling on the structure and
210 Y.L. Lyubchenko, L.S. Shlyakhtenko / Methods 47 (2009) 206–213

containing the most stable for histone octamer binding sequence


[68]. The NCPs consisting of bright blobs (nucleosome) with DNA
arms not involved into wrapping are seen all over the scan. Impor-
tantly, this sample was prepared by the deposition of the NCP sam-
ple without fixation with glutaraldehyde typically used for
electron microscopy studies, suggesting that the AFM procedure
is gentle enough to preserve the NCP intact. Quantitative analysis
of the NCP images allows one to calculate such an important
parameter of NCP’s as the number of DNA turns around the histone
cores. For NCPs marked as 1, 2 and 3 (Fig. 5) these values are 1.7,
Fig. 4. Time-lapse AFM images (6–13) illustrating unfolding (frames 6–9), dynam-
1.4 and 1 turns, respectively. Such a broad variability of the wrap-
ics of unfolded conformation (9–12) and branch migration of the Holliday junction ping value is an indirect indication of spontaneous dynamics of
leading to the dissociation of the junction. nucleosomes.
What are the pathways for such unwrapping? Does the core
particle roll along the DNA template like a wheel or do each of
dynamics of Holliday junctions (cruciforms) [67]. The results of the arms dissociate relatively independently? The models for the
this analysis showed that the geometry of cruciform is primarily dynamics of NCP have been proposed [69] but never been tested.
governed by the DNA supercoiling. DNA supercoiling shifts equilib- Time-lapse AFM would be an appropriate method in direct imaging
rium between folded and unfolded conformations of cruciform to- the chromatin dynamics and testing these models. For experiments
wards the folded one with the parallel orientation of the DNA in liquid the NCP sample in buffer containing in 10 mM Tris–HCl
helical strands rather than the antiparallel configuration typically (pH 7.5) and 4 mM MgCl2, was injected directly into flow cell with-
observed in synthetic Holliday junctions. The effect of DNA super- out drying. Images were acquired by using NanoScope IIId system
coiling is enhanced in the presence of cations. These findings pro- (Veeco/Digital Instruments, Santa Barbara, CA) operating in tap-
vide additional support for the role of DNA supercoiling in the ping mode in liquid. NP-Probes (NP-Veeco/Digital Instruments,
structure and dynamics of Holliday junctions (cruciforms). Inc.), with a spring constant 0.06 N/m and resonant frequency be-
tween 5 and 10 kHz were used. The continued scanning over the
3.2.2. Structure and dynamics of nucleosomes selected area (about 800  800 nm image size) was performed to
We have shown earlier that AP–mica is appropriate substrate follow the dynamics of NCPs. The result of one of the time-lapse
for imaging of chromatin and were able to get the filaments with experiments for a selected NCP is shown in Fig. 6A as a subset of
a clear-cut bead-on-spring morphology [25]. Concatemeric sea seven consecutive frames. As it is seen from the images, this partic-
urchin 5S rDNA array was used as a template, but the images did ular NCP, is initially quite compact with about two turns of DNA
not reveal a clear periodicity. However, the analysis showed that wrapped around nucleosome core (frames 51 and 52). The follow-
the positions of nucleosomes were not random, suggesting that ing frames [53–55] show that over time DNA spontaneously un-
the chromatin structure is dynamic. The dynamic feature can be wraps from the histone core and the particle in frame 54 has
observed with the use of the simplest chromatin system, nucleo- only about 1 turn left around the histone core. Additional unwrap-
some core particle (NCP) in which DNA is slightly longer than is ping occurs later, so the next image (frame 55) corresponds to fully
needed for wrapping around the histone core. Fig. 6 shows the unwrapped complex that is loosely bound, so the core dissociates
AFM image of the NCP system assembled on 352 bp DNA template, and leaves the area of observation (frame 57). The unwrapping
process of the NCP in each frame was analyzed quantitatively by
measuring the length of DNA arms, the angle between DNA arms
and the size of the nucleosome core. These data are shown as
graphs in Fig. 6B. The graph shows that the NCP dynamics is
accompanied by the elongation of both arms of DNA and the de-
crease of the blob size. All these measurements are consistent with
the NCP unwrapping process that follows the model with the inde-
pendent unwrapping of each arm. We did not observe rolling of the
core along the DNA template that may be a feature of the template
sequence that selected by the highest affinity for the core. Similar
study with the use of other sequence may clarify this issue.
Such extensive long-time observations were possible due to
finding the conditions at which segmental DNA mobility was low
enough to allow DNA adopt various local conformations permitting
the protein to move. However, further decreasing the surface–DNA
interaction to facilitate DNA mobility typically leads to sweeping
the molecule away from the scan area by a scanning tip. Thus,
the time-lapse imaging protocol should satisfy two conflicting con-
ditions—the molecule should be sufficiently mobile to allow obser-
vation of dynamic processes and at the same time the molecule
should remain within the observation area during the entire imag-
ing experiment. Although time-lapse AFM is very attractive feature
allowing to evaluate the system dynamics, one should keep in
mind that the requirement for the system to be bound to the sur-
face even intermittently influences the dynamics of the system.
Fig. 5. AFM image of nucleosomes in air. The image represents nucleosomes with
Therefore, time-lapse AFM experiments provide the propensity of
different amounts of DNA wrapped around the core particle. NCPs marked as 1, 2 the system to adopt various conformations, but do not allow one
and 3 have about 1.7, 1.4 and 1.0 turns of DNA wrapped around the core particle. to extract the time for such a dynamics. Recent advances in AFM
Y.L. Lyubchenko, L.S. Shlyakhtenko / Methods 47 (2009) 206–213 211

100 1000
B
80 800

Volume, nm
Arm length, nm
60 600

40 400

3
20 200

0 0
50 51 52 53 54 55 56
Frame number

Fig. 6. (A) Time-lapse AFM experiments for seven consecutive frames illustrating the dynamics of the nucleosome particle. The scan size is 150 nm. (B) The variability of arm
length (left Y-axis, black) and protein volume (right Y-axis, blue) with the imaging frame number (X-axis).

technology that have led to development of the system capable of Our AFM topographic studies of the complex formed by SfiI
scanning 1000 times faster than traditional instruments [70–73] restriction enzyme and DNA revealed a number of important struc-
alleviating the problem with sample immobilization. However, tural characteristics of this system, suggesting interesting dynamic
the limitation with the need of the sample to be bound to the sur- properties of this system. We have developed single molecule AFM
face for the AFM imaging does not eliminate the problem with force spectroscopy approach enabling one to measure the stability
evaluation of system dynamics times. of site-specific protein–DNA complexes involving the interaction of
two DNA sites performed by a specific complex [45]. The experi-
3.3. AFM force spectroscopy: stability and dynamics of synaptic DNA– mental setup is shown schematically in Fig. 7. The identical DNA
protein complexes duplexes containing SfiI recognition sites (shown as two parallel
red thick lines) are covalently attached to the AFM substrate and
AFM force spectroscopy allowing measurements of intermolecu- tip via flexible linker (zigzag green lines, section A). The synaptic
lar interaction as well as intramolecular stability is the area of AFM complex between the two duplexes is formed by SfiI when the du-
application with fast growing demand. In this mode, the AFM mea- plexes are in the proximity to each other and SfiI (blue tetrameric
sures the mechanical stability of the system positioned between the oval) binds to them (section B). The complex was formed by the
surface and AFM tip. Covalent attachment of the system at a well de- protein present in the solution. Upon retraction of the tip, the flex-
fined anchoring point is required for the force pulling experiments. ible linkers are stretched (section C) and the complex is ruptured
We have shown that functionalized surfaces open prospects for when the stretching limit is achieved (section D).
immobilizing the system in fast, efficient and reproducible fashion The stability of the synaptic complex formed by the enzyme and
[22,23,38,44,45,74–78]. The most promising is the approach with two DNA duplexes was probed in a series of approach–retraction
the use of silatranes with specific reactive groups allowing anchor- cycles during which the force–distance curved are captured. A typ-
ing the molecule via formation of covalent bonds. The example be- ical force–distance curve obtained using the experimental setup
low illustrates the immobilization approach utilizing maleimide for studying the synaptic complex schematically represented by
silatrane capable of binding thiol-containing molecules. Fig. 7A–D is shown in Fig. 8. An adhesive peak at the beginning

A B C D

Fig. 7. Schematics for the AFM force spectroscopy experiments for measuring the strength of the synaptic complex formed by two DNA duplexes and SfiI enzyme. DNA
duplexes (red parallel lines) are attached to the AFM tip and the surface via a flexible linker shown as a zigzag (A). The approach tip to the surface in the presence of the
enzyme (blue box) can lead to the formation of the synaptic complex (B). Retraction of the tip away from the surface leads to stretching of the linker (C). The complex
dissociates when the force reaches the rupture value for the complex. The protein can remain on one of the duplexes and dissociate later on. (For interpretation of the
references to colour in this figure legend, the reader is referred to the web version of this paper.)
212 Y.L. Lyubchenko, L.S. Shlyakhtenko / Methods 47 (2009) 206–213

Grants GM 062235, the National Science Foundation (0615590


and 0701892) and Nebraska Research Initiative (NRI).

References

[1] C. Bustamante, J. Vesenka, C.L. Tang, W. Rees, M. Guthold, R. Keller,


Biochemistry 31 (1992) 22–26.
[2] J. Yang, K. Takeyasu, Z. Shao, FEBS Lett. 301 (1992) 173–176.
[3] J. Vesenka, M. Guthold, C.L. Tang, D. Keller, E. Delaine, C. Bustamante,
Ultramicroscopy 42–44 (Pt. B) (1992) 1243–1249.
[4] Y.L. Lyubchenko, A.A. Gall, L.S. Shlyakhtenko, R.E. Harrington, B.L. Jacobs, P.I.
Oden, S.M. Lindsay, J. Biomol. Struct. Dyn. 10 (1992) 589–606.
[5] M.J. Allen, X.F. Dong, T.E. O’Neill, P. Yau, S.C. Kowalczykowski, J. Gatewood, R.
Balhorn, E.M. Bradbury, Biochemistry 32 (1993) 8390–8396.
[6] M. Hegner, P. Wagner, G. Semenza, FEBS Lett. 336 (1993) 452–456.
[7] J. Mou, D.M. Czajkowsky, Y. Zhang, Z. Shao, FEBS Lett. 371 (1995) 279–282.
[8] C. Bustamante, D. Erie, D. Keller, Curr. Opi. Struct. Biol. 4 (1994) 750–760.
[9] C. Bustamante, C. Rivetti, D.J. Keller, Curr. Opin. Struct. Biol. 7 (1997) 709–
716.
[10] Y.L. Lyubchenko, A.A. Gall, L.S. Shlyakhtenko, Meth. Mol. Biol. 148 (2001) 569–
Fig. 8. A typical force–distance curve obtained for the interactions between SfiI and 578.
DNA duplex: (B) the start point of the tip retraction from surface, the force curve in [11] Y. Lyubchenko, L. Shlyakhtenko, R. Harrington, P. Oden, S. Lindsay, Proc. Natl.
this region reflects the nonspecific adhesion forces between surface and the tip; (C) Acad. Sci. USA 90 (1993) 2137–2140.
[12] Y.L. Lyubchenko, P.I. Oden, D. Lampner, S.M. Lindsay, K.A. Dunker, Nucleic
the region is indicative of polymer linker stretching, this portion of the curve is
Acids Res. 21 (1993) 1117–1123.
fitted with worm-like chain approximation (solid line); (D) the abrupt jump in the
[13] Y.L. Lyubchenko, B.L. Jacobs, S.M. Lindsay, A. Stasiak, Scanning Microsc. 9
force curve is typical of specific bond rupture. (1995) 705–724 (discussion 24–27).
[14] Y.L. Lyubchenko, R.E. Blankenship, A.A. Gall, S.M. Lindsay, O. Thiemann, L.
Simpson, L.S. Shlyakhtenko, Scanning Microsc. Suppl. 10 (1996) 97–107.
[15] Y.L. Lyubchenko, S.M. Lindsay, in: A. Engel (Ed.), Procedures in Scanning Probe
Microscopy, Wiley & Sons, Ltd., New York, Toronto, 1998, pp. 493–496.
of the force curve (section B of the force curve) is accompanied by a [16] L.S. Shlyakhtenko, V.N. Potaman, R.R. Sinden, A.A. Gall, Y.L. Lyubchenko,
Nucleic Acids Res. 28 (2000) 3472–3477.
force extension curve (section C) corresponding to the stretching of [17] W.J. Tiner Sr., V.N. Potaman, R.R. Sinden, Y.L. Lyubchenko, J. Mol. Biol. 314
polymer linkers, followed by a rupture of the synaptic complexes (2001) 353–357.
event (section D). A series of control experiments allowed us to jus- [18] L.S. Shlyakhtenko, A.A. Gall, A. Filonov, Z. Cerovac, A. Lushnikov, Y.L.
Lyubchenko, Ultramicroscopy 97 (2003) 279–287.
tify this assignment [45]. The maleimide silatrane immobilization [19] L.S. Shlyakhtenko, L. Miloseska, V.N. Potaman, R.R. Sinden, Y.L. Lyubchenko,
strategy allowed us to determine the dissociation rate (koff) with Ultramicroscopy 97 (2003) 263–270.
a high accuracy due to the possibility to acquire a large set of the [20] Y.L. Lyubchenko, Cell. Biochem. Biophys. 41 (2004) 75–98.
[21] J.W. Pavlicek, E.A. Oussatcheva, R.R. Sinden, V.N. Potaman, O.F. Sankey, Y.L.
data. For example, the DFS experiments performed with two differ-
Lyubchenko, Biochemistry 43 (2004) 10664–10668.
ent sequences within the middle part of the DNA recognition re- [22] Y.L. Lyubchenko, S. Sherman, L.S. Shlyakhtenko, V.N. Uversky, J. Cell. Biochem.
gion revealed the sequence-dependent difference in the stability 99 (2006) 53–70.
of synaptic complex that correlated with the effect of the sequence [23] C.K. Riener, C.M. Stroh, A. Ebner, A.A. Gall, C. Klampfl, C. Romanin, Y.L.
Lyubchenko, P. Hinterdorfer, H.J. Gruber, Analytica Chimica Acta 479 (2003)
on the SfiI enzymatic activity [79]. Note again that the key in these 59–75.
experiments was the high yield of rupture events rupture events [24] Y.L. Lyubchenko, L.S. Shlyakhtenko, V.P. Potaman, R.R. Sinden, Microsc.
ca. 10–20% vs. ca. 1% for multistep procedures, e.g., [80]. Microanal. 8 (2002) 170–171.
[25] J.G. Yodh, N. Woodbury, L.S. Shlyakhtenko, Y.L. Lyubchenko, D. Lohr,
Biochemistry 41 (2002) 3565–3574.
[26] A.Y. Lushnikov, B.A. Brown 2nd, E.A. Oussatcheva, V.N. Potaman, R.R. Sinden,
4. Conclusions Y.L. Lyubchenko, Nucleic Acids Res. 32 (2004) 4704–4712.
[27] I. Lonskaya, V.N. Potaman, L.S. Shlyakhtenko, E.A. Oussatcheva, Y.L.
The preparation of the AFM substrates and probes with various Lyubchenko, V.A. Soldatenkov, J. Biol. Chem. 280 (2005) 17076–17083.
[28] A.Y. Lushnikov, V.N. Potaman, Y.L. Lyubchenko, Nucleic Acids Res. 34 (2006)
surface characteristics is a fundamental issue for reliable and rou-
e111.
tine use of AFM for studies nucleic acids and their complexes with [29] A.Y. Lushnikov, V.N. Potaman, E.A. Oussatcheva, R.R. Sinden, Y.L. Lyubchenko,
protein. In this paper, we illustrated the capabilities of the ap- Biochemistry 45 (2006) 152–158.
[30] L.S. Shlyakhtenko, J. Gilmore, A. Portillo, G. Tamulaitis, V. Siksnys, Y.L.
proach utilizing silane and recently developed silatrane surface
Lyubchenko, Biochemistry 46 (2007) 11128–11136.
chemistries. The latter was essential in facilitating the use of [31] R. Liu, B. Yuan, S. Emadi, A. Zameer, P. Schulz, C. McAllister, Y. Lyubchenko, G.
AFM in both imaging and force spectroscopy modes. Strong poly- Goud, M.R. Sierks, Biochemistry 43 (2004) 6959–6967.
anionic properties of nucleic acids were central to the development [32] R. Liu, C. McAllister, Y. Lyubchenko, M.R. Sierks, Biochemistry 43 (2004) 9999–
10007.
of reliable methods of attaching them to substrates, which is nec- [33] R. Liu, C. McAllister, Y. Lyubchenko, M.R. Sierks, J. Neurosci. Res. 75 (2004)
essary for high-resolution AFM imaging. However, attachment of 162–171.
the molecule, for example, a protein interaction with DNA at se- [34] S. Emadi, R. Liu, B. Yuan, P. Schulz, C. McAllister, Y. Lyubchenko, A. Messer, M.R.
Sierks, Biochemistry 43 (2004) 2871–2878.
lected points may require for such applications as time-lapse imag- [35] G. Yamin, L.A. Munishkina, M.A. Karymov, Y.L. Lyubchenko, V.N. Uversky, A.L.
ing. We believe that further progress in the silatrane surface Fink, Biochemistry 44 (2005) 9096–9107.
chemistry allowing controlled immobilization of specific groups [36] V.N. Uversky, G. Yamin, L.A. Munishkina, M.A. Karymov, I.S. Millett, S. Doniach,
Y.L. Lyubchenko, A.L. Fink, Brain Res. Mol. Brain Res. 134 (2005) 84–102.
of the proteins with minimal interfering with their activity will fur- [37] P.R. Dahlgren, M.A. Karymov, J. Bankston, T. Holden, P. Thumfort, V.M. Ingram,
ther facilitate such applications as time-lapse AFM imaging in the Y.L. Lyubchenko, Dis. Mon. 51 (2005) 374–385.
fast scanning mode and AFM force spectroscopy applications. [38] V.N. Uversky, A.V. Kabanov, Y.L. Lyubchenko, J. Proteome Res. 5 (2006) 2505–
2522.
[39] L.S. Shlyakhtenko, B. Yuan, S. Emadi, Y.L. Lyubchenko, M.R. Sierks,
Acknowledgments Nanomedicine 3 (2007) 192–197.
[40] H.G. Hansma, J.H. Hoh, Annu. Rev. Biophys. Biomol. Struct. 23 (1994) 115–139.
[41] Y.L. Lyubchenko, B.L. Jacobs, S.M. Lindsay, A. Stasiak, Scanning Microsc. 9
We thank A. Gall for providing us with a detailed protocol for (1995) 705–724 (discussion 24–27).
APS synthesis and A. Portillo for proofreading of the manuscript. [42] Y.L. Lyubchenko, L.S. Shlyakhtenko, Proc. Natl. Acad. Sci. USA 94 (1997) 496–
This work was supported in part by National Institutes of Health 501.
Y.L. Lyubchenko, L.S. Shlyakhtenko / Methods 47 (2009) 206–213 213

[43] A.V. Kransnoslobodtsev, L.S. Shlyakhtenko, E. Ukraintsev, T.O. Zaikova, [63] M. Guthold, X. Zhu, C. Rivetti, G. Yang, N.H. Thomson, S. Kasas, H.G. Hansma, B.
J.F.W. Keana, Y.L. Lyubchenko, Nanomed. Nanotechnol. Biol. Med. 1 (2005) Smith, P.K. Hansma, C. Bustamante, Biophys. J. 77 (1999) 2284–2294.
300–305. [64] L.S. Shlyakhtenko, E. Appella, R.E. Harrington, I. Kutyavin, Y.L. Lyubchenko, J.
[44] M.A. Karymov, A.V. Krasnoslobodtsev, Y.L. Lyubchenko, Biophys. J. 92 (2007) Biomol. Struct. Dyn. 12 (1994) 131–143.
3241–3250. [65] L.S. Shlyakhtenko, D. Rekesh, S.M. Lindsay, I. Kutyavin, E. Appella, R.E.
[45] A.V. Krasnoslobodtsev, L.S. Shlyakhtenko, Y.L. Lyubchenko, J. Mol. Biol. 365 Harrington, Y.L. Lyubchenko, J. Biomol. Struct. Dyn. 11 (1994) 1175–1189.
(2007) 1407–1418. [66] A.Y. Lushnikov, A. Bogdanov, Y.L. Lyubchenko, J. Biol. Chem. 278 (2003)
[46] L.S. Shlyakhtenko, V.N. Potaman, R.R. Sinden, Y.L. Lyubchenko, J. Mol. Biol. 280 43130–43134.
(1998) 61–72. [67] A.L. Mikheikin, A.Y. Lushnikov, Y.L. Lyubchenko, Biochemistry 45 (2006)
[47] A. Engel, Y. Lyubchenko, D. Muller, Trends Cell Biol. 9 (1999) 77–80. 12998–13006.
[48] L.S. Shlyakhtenko, P. Hsieh, M. Grigoriev, V.N. Potaman, R.R. Sinden, Y.L. [68] K.J. Polach, J. Widom, J. Mol. Biol. 254 (1995) 130–149.
Lyubchenko, J. Mol. Biol. 296 (2000) 1169–1173. [69] J. Widom, Q. Rev. Biophys. 34 (2001) 269–324.
[49] T. Thundat, D.P. Allison, R.J. Warmack, G.M. Brown, K.B. Jacobson, J.J. Schrick, [70] T. Ando, N. Kodera, E. Takai, D. Maruyama, K. Saito, A. Toda, Proc. Natl. Acad.
T.L. Ferrell, Scanning Microsc. 6 (1992) 911–918. Sci. USA 98 (2001) 12468–12472.
[50] H.G. Hansma, D.E. Laney, M. Bezanilla, R.L. Sinsheimer, P.K. Hansma, Biophys. J. [71] M. Kobayashi, K. Sumitomo, K. Torimitsu, Ultramicroscopy 107 (2007) 184–
68 (1995) 1672–1677. 190.
[51] C. Pfannschmidt, A. Schaper, G. Heim, T.M. Jovin, J. Langowski, Nucleic Acids [72] T. Ando, T. Uchihashi, N. Kodera, D. Yamamoto, A. Miyagi, M. Taniguchi, H.
Res. 24 (1996) 1702–1709. Yamashita, Pflugers Arch. 456 (2008) 211–225.
[52] K. Rippe, N. Mucke, J. Langowski, Nucleic Acids Res. 25 (1997) 1736–1744. [73] N. Crampton, M. Yokokawa, D.T. Dryden, J.M. Edwardson, D.N. Rao, K.
[53] C. Pfannschmidt, J. Langowski, J. Mol. Biol. 275 (1998) 601–611. Takeyasu, S.H. Yoshimura, R.M. Henderson, Proc. Natl. Acad. Sci. USA 104
[54] S.D. Jett, D.I. Cherny, V. Subramaniam, T.M. Jovin, J. Mol. Biol. 299 (2000) 585– (2007) 12755–12760.
592. [74] A.P. Limansky, L.S. Shlyakhtenko, S. Schaus, E. Henderson, Y.L. Lyubchenko,
[55] F. Nagami, G. Zuccheri, B. Samori, R. Kuroda, Anal. Biochem. 300 (2002) 170– Probe Microsc. 4 (2002) 1–6.
176. [75] A.V. Kransnoslobodtsev, L.S. Shlyakhtenko, E. Ukraintsev, T.O. Zaikova, J.F.
[56] A. Vologodskii, N.R. Cozzarelli, Biophys. J. 70 (1996) 2548–2556. Keana, Y.L. Lyubchenko, Nanomedicine 1 (2005) 300–305.
[57] V.V. Rybenkov, A.V. Vologodskii, N.R. Cozzarelli, Nucleic Acids Res. 25 (1997) [76] C. McAllister, M.A. Karymov, Y. Kawano, A.Y. Lushnikov, A. Mikheikin, V.N.
1412–1418. Uversky, Y.L. Lyubchenko, J. Mol. Biol. 354 (2005) 1028–1042.
[58] S.Y. Shaw, J.C. Wang, Science 260 (1993) 533–536. [77] C. Wei, Y.L. Lyubchenko, H. Ghandehari, J. Hanes, K.J. Stebe, H.-Q. Mao, D.T.
[59] V.N. Potaman, L.S. Shlyakhtenko, E.A. Oussatcheva, Y.L. Lyubchenko, V.A. Haynie, D.A. Tomalia, M. Foldvari, N. Monteiro-Riviere, P. Simeonova, S. Nie, H.
Soldatenkov, J. Mol. Biol. 348 (2005) 609–615. Mori, S.P. Gilbert, D. Needham, Nanomed. Nanotechnol. Biol. Med. 2 (2006)
[60] B. Drake, C.B. Prater, A.L. Weisenhorn, S.A. Gould, T.R. Albrecht, C.F. Quate, D.S. 253–263.
Cannell, H.G. Hansma, P.K. Hansma, Science 243 (1989) 1586–1589. [78] C. Wei, N. Liu, P. Xu, M. Heller, D.A. Tomalia, D.T. Haynie, E.H. Chang, K. Wang,
[61] M. Guthold, M. Bezanilla, D.A. Erie, B. Jenkins, H.G. Hansma, C. Bustamante, Y.S. Lee, Y.L. Lyubchenko, R. Bawa, R. Tian, J. Hanes, S. Pun, J.C. Meiners, P. Guo,
Proc. Natl. Acad. Sci. USA 91 (1994) 12927–12931. Nanomedicine 3 (2007) 322–331.
[62] C. Bustamante, C. Rivetti, Annu. Rev. Biophys. Biomol. Struct. 25 (1996) [79] S.A. Williams, S.E. Halford, Nucleic Acids Res. 29 (2001) 1476–1483.
395–429. [80] C. Ray, B.B. Akhremitchev, J. Am. Chem. Soc. 127 (2005) 14739–14744.

Anda mungkin juga menyukai