Methods in Enzymology Chromatin and Chromatin Remodeling Enzymes
Methods in Enzymology Chromatin and Chromatin Remodeling Enzymes
Methods in Enzymology Chromatin and Chromatin Remodeling Enzymes
C. David Allis
Carl Wu
FOUNDING EDITORS
Martin L. Bennink (6), Biophysical Tech- Peter Cheung (15), Department of Med-
niques Group and MESAþ Research ical Biophysics, University of Toronto,
Institute, Department of Science Technol- Ontario Cancer Institute, Toronto,
ogy, University of Twente, 7500 AE Ontario M5G 2M9, Canada
Enschede, The Netherlands
J. Chin (1), Department of Biochemistry,
Bradley E. Bernstein (23), Department of Northwestern University, Molecular Bio-
Chemistry and Chemical Biology, Harvard logy and Cell Biology, Evanston, Illinois
University, Cambridge, Massachusetts 60208-3500
02138
David N. Ciccone (22), Department of
Margie T. Borra (11), Department of Bio- Molecular Biology, Massachusetts Gen-
chemistry and Molecular Biology, Oregon eral Hospital, Boston, Massachusetts
Health and Science University, Portland, 02114
Oregon 97239
Philip A. Cole (12), Department of
Brent Brower-Toland (5), Biology De- Pharmacology and Molecular Sciences,
partment, Washington University in St. Johns Hopkins University School of
Louis, St. Louis, Missouri 63130 Medicine, Baltimore, Maryland 21218
Michael Bustin (14), Protein Section, Carlos Cordon-Cardo (13), Division of
National Cancer Institute, National Insti- Molecular Pathology, Memorial Sloan
tutes of Health, Bethesda, Maryland Kettering Cancer Center, New York,
20892 New York 10021
ix
x contributors to volume 376
Peggy Farnham (21), McArdle Laboratory Monika Kauer (16), Research Institute of
for Cancer Research, University of Wis- Molecular Pathology (IMP), The Vienna
consin, Madison, Wisconsin 53706 Biocenter, Vienna, A-1030, Austria
Victoria M. Richon (13), Discovery Biol- Ling Wang (12), Department of Pharma-
ogy, Aton Pharma, Inc., Tarrytown, New cology and Molecular Sciences, Johns
York 10591 Hopkins University School of Medicine,
Baltimore, Maryland 21218
Richard C. Robinson (14), Laboratory of
Metabolism, National Institutes of Janis K. Werner (26), Cornell University,
Health, National Cancer Institute, Ithaca, New York 14853
Bethesda, Maryland 20892 Jon Widom (1), Northwestern University,
Daniel Robyr (19), Department of Biol- Department of Biochemistry, Molecular
ogical Chemistry, University of California, Biology and Cell Biology, Evanston,
Los Angeles, School of Medicine and Illinois 60208-3500
Molecular Biology Institute, Los Angeles, Christopher L. Woodcock (3), Department
California 90095 of Biology, University of Massachusetts,
Kavitha Sarma (17), Department of Biol- Amherst, Massachusetts 01003
ogy, Howard Hughes Medical Institute,
Patrick Yau (21), Microarray Centre, Uni-
University of Medicine and Dentistry
versity Health Network, Toronto, Ontario
of New Jersey, Piscataway, NJ
M5G 2C4, Canada
08854-5635
Ken Zaret (24), Cell and Developmental
Stuart Schreiber (23), Department of
Biology Program, W. W. Smith Chair in
Chemistry and Chemical Biology, Harvard
Cancer Research, Fox Chase Cancer
University, Cambridge, Massachusetts
Center, Philadelphia, Pennsylvania 19111
02138
Yujun Zheng (12), Department of
BrianE.Schwartz(26),CornellUniversity,
Pharmacology and Molecular Sciences,
Ithaca, New York 14853
Johns Hopkins University School of
J. Paul Secrist (13), Discovery Biology, Medicine, Baltimore, Maryland 21218
Aton Pharma, Inc., Tarrytown, New York
10591 Ming-Ming Zhou (8), Structural Biology
Program, Department of Physiology and
Gena E. Stephens (25), Biology Depart- Biophysics, Mt. Sinai School of Medicine,
ment, Washington University in St. Louis, New York University, New York, New
St. Louis, Missouri 63130 York 10029-6574
contributors to volume 376 xiii
Xianbo Zhou (13), Discovery Biology, Aton Jordanka Zlatanova (6), Department of
Pharma, Inc., Tarrytown, New York Chemical and Biological Sciences and En-
10591 gineering, Polytechnic University, Brook-
lyn, New York 11201
[1] fluorescence anisotropy assays 3
1
T. Tsukiyama, C. Daniel, J. Tamkun, and C. Wu, Cell 83, 1021 (1995).
2
P. D. Varga-Weisz et al., Nature 388, 598 (1997).
3
G. Längst and P. B. Becker, J. Cell Sci. 114, 2561 (2001).
set up with the probe concentration Kd; consequently the free concentra-
tion of the added macromolecule (ISWI, in our case), which is generally
either difficult to measure or is completely unknown, will be approximately
equal to the total concentration, which can be definitively measured, thus
greatly simplifying the analysis of the binding measurements. Another im-
portant benefit of the sensitivity of the anisotropy measurement is that it
preserves precious reagents. Measurements can be made in small volumes,
and samples can be recovered and reused if desired.
Finally, as discussed later, the experiment can be carried out using inex-
pensive conventional fluorometers such as are found at most biochemical
or chemical research laboratories, or, alternatively, using an inexpensive
instrument specialized for the fluorescence anisotropy experiment.
Investigators planning to carry out such studies should study two par-
ticularly useful references, one on fluorescence theory and methodology
in general4 and one focused on fluorescence approaches to analysis of pro-
tein-DNA interactions in particular.5 These references nicely define and
explain the set of four fluorescence intensity measurements that go into a
single measurement of fluorescence anisotropy; we will not duplicate this
important topic here, but rather refer readers to these other sources.
Fluorescein-Labeled DNA
We use DNA sequences labeled with fluorescein attached at the 50 -end
through a C6 linker. Relatively short sequences are purchased as a pair
of complementary oligonucleotides, one containing 50 -fluorescein. These
are annealed, and the resulting duplex purified away from any remaining
single strand by reverse-phase HPLC on a Zorbax-10 column using a
gradient of 10–20% acetonitrille in 0.1 M triethanolamine-acetate, pH
7.0, 0.1 mM EDTA, developed over 10 min at 1 ml min1. When longer
sequences (e.g., nucleosome-length DNAs) are required, direct synthesis
is not practical. Instead we use preparative scale PCR, with one of
the two primers again containing 50 -fluorescein. The resulting PCR
product is purified by gel electrophoresis in 1% agarose gels with standard
TAE buffer, and extracted from the gel using Ultra-DA (Millipore) gel
extraction kits.
DNA concentrations are quantified by UV absorbance.
4
J. R. Lakowicz, ‘‘Principles of Fluorescence Spectroscopy,’’ 2nd Ed. Kluwer Academic/
Plenum Press, New York, 1999.
5
J. J. Hill and C. A. Royer, Meth. Enzymol. 278, 390 (1997).
[1] fluorescence anisotropy assays 5
Preparation of Nucleosomes
Nucleosomes are formed by salt gradient dialysis using purified histone
octamer and DNA, and the resulting nucleosomes are purified by sucrose
gradient ultracentrifugation, as described.6–8 We typically label a small
amount of the fluorescein-labeled DNA with [ 32] ATP (at the 50 end that
does not have a fluorescein) using T4 polynucleotide kinase to facilitate
following the sample throughout preparation and purification. Reconstitu-
tion reactions typically contain 300 ng of (32P, fluorescein) double-labeled
DNA, 15 g of fluorescein-only labeled DNA, 3 g of histone octamer, in a
300 l volume of 2.5 M NaCl, 0.5 TE (TE is 10 mM Tris, pH 8.0, 1 mM
Na3 EDTA) with 0.5 mM phenylmethylsulfonyl fluoride (PMSF), and
0.1 mM benzamidine (BZA) added as protease inhibitors. The reconstitu-
tion reactions are loaded onto 12 ml 5–30% sucrose gradients in 0.5 TE
and centrifuged at 4 in an SW41 rotor (Beckman) at 41,000 rpm for
22–24 h. (We aim for substoichiometric reconstitution of histone octamer
onto DNA, as this eliminates the possibility of overloading nucleosomes
with excess histones9 while providing useful diagnostics for the reconstitu-
tion and markers for the subsequent sucrose gradient purification.) Gradi-
ents are fractionated from the bottom in 0.5-ml fractions; fractions
containing nucleosomes are identified by scintillation counting, pooled
and exchanged into 0.5 TE buffer on Centricon-30 concentrators, and
analyzed by native polyacrylamide gel electrophoresis. Nucleosome con-
centrations are measured by UV absorbance at 260 nm. Reconstituted
nucleosomes are stored at concentrations of 50 nM or greater, on ice in
0.5 TE, and are used within 2 weeks.
6
K. J. Polach and J. Widom, J. Mol. Biol. 254, 130 (1995).
7
P. T. Lowary and J. Widom, J. Mol. Biol. 276, 19 (1998).
8
J. D. Anderson, A. Thåström, and J. Widom, Mol. Cell. Biol. 22, 7147 (2002).
9
G. Voordouw and H. Eisenberg, Nature 273, 446 (1978).
6 chromatin proteins [1]
Sample Cleanliness
As with any sensitive experimental method in analytical biochemistry,
care must be taken in certain matters to avoid potential pitfalls.
It goes without saying that both buffers and samples must be free from
significant fluorescent contaminants. Fluorescence from the buffer alone
and from unlabeled samples should be checked and shown to be negligible
in comparison to the fluorescence obtained at the desired concentration of
labeled sample. In our experience this has never proven to be a problem
using dilute buffers supplemented with approximately physiological con-
centrations of salts and Mg2þ and small amounts of glycerol; nevertheless,
it should be checked, especially in case of problems with the water or with
contaminated glass- or plastic-ware.
Scattered Light
Even when samples are free of contaminants, one must take care to
eliminate certain additional potential artifacts due to scattered light. Scat-
tered light is particularly problematic in anisotropy measurements because
it is generally perfectly polarized, and hence will systematically distort
measurements of fluorescence anisotropy from the sample.
Two chief types of scattered light need to be considered in anisotropy
measurements: elastic (Rayleigh) scattering and inelastic (chiefly Raman)
scattering. Elastic scattering is a process in which excitation light is scat-
tered in all directions, unshifted in wavelength, by interaction of the excita-
tion light with molecules in the sample. Even pure solvents scatter light
elastically. Such scattering is weak, yet may nevertheless be significant
in comparison to the faint fluorescence from a very dilute fluorophore.
Macromolecules in solution greatly increase the intensity of scattered light,
in proportion to their concentration and molecular weight. Solutions con-
taining a high molecular weight species such as nucleosomes can result in
a scattering intensity that greatly exceeds the intensity of fluorescence
emission from a dye attached to that same macromolecule.
If excitation and emission monochrometers were ‘‘perfect,’’ then elastic
scattering would present no problem: one could simply set the excitation and
emission monochrometers to different wavelengths (e.g., the excitation
and emission maxima, respectively), and there would be no leakage of scat-
tered excitation light through the emission monochrometer. In fact, how-
ever, the finite resolution of the monochrometers, together with optical
imperfections that allow light of colors outside the assumed bandpass to
pass through, albeit at reduced intensity, are such that there can be signifi-
cant excitation intensity at the color chosen for emission measurement,
even though these colors may differ by 30 nm or more. In fact, this leakage
[1] fluorescence anisotropy assays 7
fluorescein. The colored glass cut-on filter rejects most of the excitation
bandpass, while passing the majority of the emission spectrum. The special-
ized cut-off filter nicely rejects any Raman scatter, which is centered
at 595 nm, while further strongly reducing any elastically scattered excita-
tion light. (The cut-off [bandpass] filter also strongly rejects elastically
scattered light on its own; however, the combination of the two filters gives
far better blocking at the excitation bandpass than either one on its own.
This improved scatter rejection is important in strongly scattering/weakly
fluorescing samples.)
The performance of this filter set can be appreciated from the emission
spectra in Fig. 2. Panel A shows the ability of the filter set to suppress both
elastically scattered light (>10,000-fold reduction) and Raman scattering to
undetectably low levels, while panel B shows that this huge reduction in
background scattering intensity comes at a modest cost (2- to 3-fold) in
the collectible intensity of fluorescence emission.
Fig. 2. Performance of the filter set for fluorescein anisotropy measurement. (A) Rejection
of elastically scattered light and Raman scatter. Emission spectra recorded from a scattering
solution (10 g/ml BSA in buffer), with excitation at 490 nm using the excitation mono-
chrometer plus the 488.8 nm interference filter. Solid line: no filters in the emission path. Long-
dashed line: 515 nm cut-on filter. Short-dashed line (essentially invisible): both 515-nm cut-on
plus bandpass filters. When no filters are used in the emission path, note the strong peak of
scattering intensity detected when emission ¼ excitation. Inset: same spectra plotted on a
500-fold more sensitive scale. The emission filter combination reduces the scattering signal by
>10,000-fold, rendering it immeasurably low. Similarly, there is negligible remaining intensity
at the Raman scattering wavelength (595 nm). (B) Fluorescence emission from fluorescein
with no filter in the emission path (solid black line), 515-nm cut-on filter only (grey line) or
both cut-on and bandpass filters (dashed line). The >10,000-fold reduction in scattering
intensity comes at a cost of only 2-fold in emission intensity at the emission maximum, and
only 3-fold in total emission intensity integrated over all wavelengths (such as could be
measured by a detector in the emission path if no emission monochrometer were present).
[1] fluorescence anisotropy assays 11
pH Sensitivity
Fluorescein has a pKa of 6.5, and its fluorescence intensity is depen-
dent on the charge state. Consequently, it is important to control the pH
of the solution, preferably at a pH 7.5, so that unanticipated small
changes in pH will affect the fluorescence intensity only negligibly.
Intensity Changes
Two additional phenomena concerning the sample fluorescence inten-
sity bear mention. First, fluorescein is sensitive to photobleaching; thus,
prolonged exposure of sample to excitation light, or even room light, can
significantly reduce the fluorescence intensity. Since anisotropy is an inher-
ent property, independent of concentration, modest amounts of sample
bleaching occurring during binding titrations need not be problematic, pro-
vided that there is negligible bleaching during the time required for any
given anisotropy measurement. This can be checked by monitoring the
total intensity during the four individual measurements that make up an
anisotropy measurement.
In certain cases, however, the binding of a protein to a fluorescein-
labeled DNA can affect the fluorescence intensity directly. For example,
the bound protein might happen to quench (or enhance) the fluorescence
quantum yield, or perhaps shift the emission color, so that the intensity
measured over a given wavelength range may increase or decrease. Such
effects invalidate the usual interpretation of the anisotropy changes. If
the intensity changes, the likelihood is that the fluorescence lifetime too
is changing; and if the lifetime changes, then the anisotropy is affected even
if molecular tumbling time were to remain constant. Consequently, it will
no longer be possible to assign a linear relationship between a measured
anisotropy change and a probability of binding site occupancy.
Actually, such cases can be a blessing in disguise: one can often monitor
the binding process simply by measuring the intensity change directly. In
any event, it is necessary to pay attention to the total intensity when carry-
ing out anisotropy measurements. Systematic changes in intensity that cor-
relate with binding invalidate the standard interpretation of the anisotropy
experiment.
Cuvettes
Samples used in fluorescence experiments may be precious, and one
may wish to minimize the volume of sample used. We routinely use
samples of 75–100 l in quartz ultramicro-cuvettes (Hellma, black walled,
#105.251-QS) having a sample chamber of 3 3 mm with a 5-mm tall
12 chromatin proteins [1]
aperture, that is, a 45-l illuminated volume. These cuvettes are easy to fill
and clean, and fit in the 1.25-cm square cuvette holders that are standard in
most fluorometers. These cuvettes are available with the sample chamber
placed at various heights above the base of the cuvette. It is important to
pay attention to the height of the optical axis of the fluorometer, and to
make certain that the height of the sample chamber of the cuvette matches
that of the fluorometer.
10
A. Thåström et al., J. Mol. Biol. 288, 213 (1999).
11
J. Widom, Q. Rev. Biophys. 34, 269 (2001).
12
G. Längst and P. B. Becker, Mol. Cell 8, 1085 (2001).
13
D. F. Corona et al., Mol. Cell 3, 239 (1999).
[1] fluorescence anisotropy assays 13
Fig. 3. Raw fluorescence anisotropy data from ISWI binding to naked DNA. Titration of a
5 nM 50 -fluorescein–labeled 35-bp long DNA with increasing concentrations of ISWI protein.
(See text for buffer conditions.) The curve superimposed on the data represents a fit to a
cooperative binding model.
combination of the cut-on and bandpass cut-off filter in the emission path
described previously.
We always include a sample prepared with no ISWI (ISWI ¼ 0) to es-
tablish the experimental ‘‘baseline’’ for each titration, and we extend the
titrations to sufficiently high ISWI to allow accurate determination of the
anisotropy corresponding to complete binding (complete occupancy of
binding sites by bound ISWI). Note that, whatever method is used to moni-
tor binding processes, it is important to carry titrations through the full
range of the binding process. A good practice is to use ‘‘direct’’ plots14 of
the measured signal (anisotropy, in our case) versus the titrant concentra-
tion (ISWI, in our case) plotted on a log scale to allow representation of the
wide range of titrant necessary to explore the full range from fraction
bound 0 (no binding) to fraction bound 1 (100% binding).
We use Kalaidagraph software to fit raw binding data to desired binding
models.
14
I. M. Klotz, ‘‘Ligand-Receptor Energetics: A Guide for the Perplexed.’’ Wiley, New York,
1997.
14 chromatin proteins [1]
affinity of ISWI for naked DNA is not influenced by the presence of high
concentrations of ADP. This figure highlights the utility of anisotropy
measurements to monitor binding in solutions containing other additives
such as nucleotides. In contrast, it is difficult or risky to carry out such stud-
ies using a gel electrophoretic mobility shift approach, for example, since
prohibitively costly amounts of nucleotide analogs might be required, and
moreover these compounds may actually electrophorese. In that case, their
concentrations around the complexes during a gel separation would be
undefined, rendering the experiments uninterpretable.
Conclusions
Fluorescence anisotropy is well known to be useful for analysis of pro-
tein-DNA interactions, and it seems likely to be particularly useful for
analysis of nucleosome remodeling factors because it is rapid, quantitative,
highly sensitive (conserving precious reagents), suitable for use in the pres-
ence of cofactors such as ATP, readily measured even during rapid kinetic
experiments, and very broadly applicable. It will allow analysis of the inter-
actions of remodeling factors or their individual proteins or domains with
any other species that can be specifically labeled.
[2] biophysical analysis of genomic loci 17
Background
Over the last decade much progress has been made toward understand-
ing the effects of chromatin on nuclear functions. However, virtually all
previous studies of chromatin fiber organization in vivo have been re-
stricted to gathering information about the locations of nucleosomes, his-
tone post-translational modifications, regulatory DNA binding proteins,
and chromatin remodeling machines relative to specific functional DNA
elements, for example, promoters, origins of replications, repair sites.
Despite a vastly improved understanding of the composition and configur-
ation of functionally important genomic loci, very little is known about the
higher-order organization of these chromosomal regions in vivo. Biophys-
ical characterization of specific in vivo–assembled chromatin structures has
not been possible due to technical limitations. Consequently, our knowl-
edge of the functional effects of chromatin folding and higher-order struc-
ture has been obtained almost exclusively through use of in vitro model
systems that mimic the solution behavior of natural chromatin.1–4
Recently, we have adapted the technique of agarose multigel electro-
phoresis (AME)5–8 for analysis of the higher-order nucleoprotein structure
of specific genomic loci that have been isolated as native chromatin in
unfractionated low-salt nuclear extracts.9,10 This approach yields analytical
measurements of average macromolecular radii and surface charge density,
which in turn allows one to evaluate the condensation behavior and
conformational flexibility of the chromatin fragment being studied. Here
1
J. C. Hansen and C. L. Turgeon, Methods Mol. Biol. 119, 127 (1999).
2
J. C. Hansen, Annu. Rev. Biophys. Biomol. Struct. 31, 361 (2002).
3
P. J. Horn, K. A. Crowley, L. M. Carruthers, J. C. Hansen, and C. L. Peterson, Nat. Struct.
Biol. 9, 167 (2002).
4
B. Dorigo, T. Schalch, K. Bystricky, and T. J. Richmond, J. Mol. Biol. 327, 85 (2003).
5
T. M. Fletcher, P. Serwer, and J. C. Hansen, Biochemistry 33, 10859 (1994).
6
T. M. Fletcher, U. Krishnan, P. Serwer, and J. C. Hansen, Biochemistry 33, 2226 (1994).
7
L. M. Carruthers, C. Tse, K. P. Walker, III, and J. C. Hansen, Methods Enzymol. 304,
19 (1999).
8
L. M. Carruthers and J. C. Hansen, J. Biol. Chem. 275, 37285 (2000).
9
P. T. Georgel and J. C. Hansen, Biopolymers 68, 557 (2003).
10
P. T. Georgel, T. M. Fletcher, G. L. Hager, and J. C. Hansen, Genes Dev. 17, 1617 (2003).
11
C. L. Woodcock and S. Dimitrov, Curr. Opin. Genet. Dev. 11, 130 (2001).
12
G. A. Griess, E. T. Moreno, R. A. Easom, and P. Serwer, Biopolymers 28, 1475 (1989).
13
J. C. Hansen, J. I. Kreider, B. Demeler, and T. M. Fletcher, Methods 12, 62 (1997).
[2] biophysical analysis of genomic loci 19
Fig. 1. (A) Multigel apparatus. (B) 9-lane agarose multigel. Percentages of agarose in
running gels are in increment of 0.1%. Note: the figure shows only one half of an 18-lane gel
(mirror image).
In an AME experiment, the sample of interest is first spiked with the spher-
ical bacteriophage T3 (Re ¼ 30.1 nm). The T3 00 is obtained by extrapolat-
ing the linear region of a plot of log T3 versus agarose percentage to 0%
agarose, that is, the y-axis.5,6,13 Using Eq. (1), the Pe of each running gel is
calculated from the and 00 and known Re (30.1 nm) of the T3 internal
standard. For the unknown band(s) of interest, the and 00 are determined
experimentally and the Re in each running gel is calculated using Eq. (1)
and the T3-derived values of Pe for that gel.
20 chromatin proteins [2]
14
J. C. Hansen, J. Ausio, V. H. Stanik, and K. E. van Holde, Biochemistry 28, 9129 (1989).
15
J. D. Dignam, R. M. Lebovitz, and R. G. Roeder, Nucleic Acids Res. 11, 1475 (1983).
[2] biophysical analysis of genomic loci 21
Methods
The only aspect of the AME approach that is not general is obtaining
the specific genomic chromatin fragment of interest in a low-salt nuclear
extract. Once this has occurred, the remainder of the method has been
standardized. Presented in the following are the general protocols needed
to:
1. Prepare a multigel.
2. Perform an AME experiment using low-salt nuclear extracts as the
source of the chromatin.
3. Transfer the multigel to membrane and measure the in each
running gel through use of Southern blotting.
4. Calculate the , 00 , and Re of the chromatin-band(s) of interest, and
the Pe of each running gel.
In addition, we describe western assays for determining the compo-
sition of the genomic chromatin fragment, and biophysical assays for
formation of salt-dependent secondary chromatin structure and for the
extent of conformational flexibility. To emphasize the potential of the
AME approach, we conclude the article by briefly describing key aspects
of our recent use of AME to characterize the active and inactive states of
genomic MMTV promoters.
Equipment
Multigel units such as those shown in Fig. 1A are commercially avail-
able from Aquebogue (NY). It is important to keep all the various com-
ponents of the multigel together as a single unit. A high-temperature
water bath is needed to prepare the multigel. A Cole Parmer oscillating
pump run at 40–50% maximum output voltage (i.e., 50–60 V for this pump)
regulated through the use of a rheostat is needed to circulate the running
buffer and maintain temperature control. For the reasons described later,
we strongly recommend that the agarose be molecular biology grade,
Low Electro Endo-Osmosis purchased from Research Organics (cat #
1170A-3). The casting and running buffer is either E buffer (40 mM Tris-
HCl [pH 7.8], 0.25 mM Na2 EDTA) or TAE (40 mM Tris acetate [pH
8.3], 1 mM Na2EDTA). When formation of secondary chromatin struc-
tures is being assayed, E buffer is used and contains MgCl2 at concentra-
tions equal to 0.1–2.0 mM free Mg2þ. Do not use Tris-borate buffers in
combination with low electro endo-osmosis (EEO) agarose, as this will
generate anomalous undesirable electro-osmosis effects.
22 chromatin proteins [2]
Multigel Preparation
Set the high-temperature water bath at 65–70 , and fill with enough
water to have the entire tube of agarose (see later) submerged to about
1 cm from the lid. The gel frame is prepared with 1.5% agarose in E or
TAE buffer. Weigh 1.8 g of agarose into a 250-ml flask. Tare the scale.
Add 120 ml of running buffer (measure by weight not by volume; i.e., as-
suming a buffer density of 1.0, add 120 g of buffer) and tare again. Cover
the top of the flask with a piece of plastic wrap, punch a small hole, and
melt the agarose in a microwave oven. After melting, weigh the flask again.
Compensate for evaporation by adding milliQ water. Cool the agarose to a
temperature around 55 (i.e., comfortable to touch). Assemble the slot
former and the desired comb as described in detail,16 and as shown in
Fig. 1A. Secure the end plates in the grooves with the détentes (small plas-
tic parallelepipedic blocks). If needed, seal the plates with a small quantity
of 1.5% framing agarose. Level the unit. To make the frame, pour the
molten agarose into the gel bed, allow the agarose to set for 30 or 60 min
(9- or 18-lane slot formers, respectively).
Begin preparing the agarose running gels. Start by labeling capped 15-ml
Kimble borosilicate (threaded end, 20 125 for 9-lane gels and 16 125 for
18-lane gels) tubes with the chosen agarose concentrations (see Table I
for details). Add running buffer to the tubes (see Table I for volumes)
and place the tubes in the water bath at 65–70 .
After complete polymerization of the frame, carefully remove the comb
and then very gently pull out the slot former. Abrupt removal will result in
breaking the lanes. You may want to practice removing the slot former
using standard agarose prior to using the more expensive low EEO agar-
ose. Using a pair of small forceps, remove the small strips of agarose that
often form at the edges of the wells. Burst any bubbles that may have
formed in the thin layer of agarose beneath each slot. Put the comb back
in its original location.
While the frame is polymerizing, prepare the agarose stock(s) for the gel
dilutions. To prepare a 0.2–1% agarose multigel, start with 60 ml of molten
1% agarose in a 125-ml flask (prepared in the same way as described earlier
for the 1.5% agarose stock). Immediately dilute the stock according to
Table I to prepare each running gel. After adding the appropriate volume
of 1% agarose to the pre-warmed running buffer, cap the tube, vortex
briefly, and immediately pour the agarose into the appropriate multigel slot
using disposable plastic pipettes. Prepare one agarose concentration at a
time. Use a new pipette every time you change agarose concentration.
16
J. C. Hansen, T. M. Fletcher, and J. I. Kreider, Methods Mol. Biol. 119, 113 (1999).
[2] biophysical analysis of genomic loci 23
TABLE I
Dilutions
1% stock
18-lane frame
0.2 4 1
0.3 3.5 1.5
0.4 3 2
0.5 2.5 2.5
0.6 2 3
0.7 1.5 3.5
0.8 1 4
0.9 0.5 4.5
1 0 5
9-lane frame
0.2 8 2
0.3 7 3
0.4 6 4
0.5 5 5
0.6 4 6
0.7 3 7
0.8 2 8
0.9 1 6
1 0 10
3% stock
18-lane frame
0.9 3.5 1.5
1 3.4 1.7
1.3 2.9 2.2
1.6 2.4 2.7
1.9 1.9 3.2
2.2 1.4 3.7
2.5 0.9 4.2
2.8 0.4 4.7
3 0 5
9-lane frame
0.9 7 3
1 6.7 3.3
1.3 5.7 4.3
1.6 4.7 5.3
1.9 3.7 6.3
2.2 2.7 7.3
2.5 1.7 8.3
2.8 0.7 9.3
3 0 10
24 chromatin proteins [2]
For a 9-lane frame, use 7 ml of diluted agarose per lane and 2.5 ml for
an 18-lane. Note that to maintain stability of the frame, the 0.2% agarose
should not be poured in the slot at the edge of the multigel frame. For a
0.2–1.0% multigel we suggest the following pouring order: 0.8, 0.9, 1.0,
0.2, 0.3, 0.4, 0.5, 0.6, 0.7. To prepare a 1.0–3.0% agarose 18-lane multigel,
start with 60 ml of molten 3.0% agarose in a 125-ml flask. For each running
gel, immediately dilute the stock according to Table I as described previ-
ously. For a 0.9–3.0% gel we suggest the following pouring order: 0.9, 1.0,
1.3, 1.6, 1.9, 2.2, 2.5, 2.8, 3.0.
Let the running gels set for 1 h. Pull out the comb, end plates, and dé-
tentes. Fill the multigel apparatus with running buffer. The multigels can be
prepared in advance and used the following day. Make sure that the gel box
is covered with plastic wrap to avoid unnecessary evaporation of buffer.
Gels have been used successfully after 3 days. To prevent the gel from
sliding off the gel bed during subsequent electrophoresis, place several
10-l plastic tips in the plate grooves at opposite corners.
Gel Electrophoresis
Prior to electrophoresis, low-salt nuclear extracts (100 g total
nucleic acid) are transferred to fresh tubes containing the appropriate
amount of empirically determined herring sperm DNA, 0.5 g of bacterio-
phage T3, and glycerol (10% w/v final). Samples are loaded in the
appropriate well and electrophoresed at 1.33 V/cm for 6 h (timed manually
and precisely). Alternatively, one can perform the electrophoresis for
1 V/cm for 8 h.13 The average buffer temperature in our experiments
has been 24 3 (room temperature). Using the pump system described
earlier, the running buffer was continuously recirculated throughout the
entire experiment at a slow and steady pace, which prevents the formation
of pH or ion gradients and large temperature fluctuations.
To measure the of the bacteriophage T3 in each running gel,
after electrophoresis a thin plastic sheet is placed under the multigel
to permit transfer of the fragile gel to a tray containing either ethidium
bromide or SYBR green. The tray is placed on a rotating shaker set at
60 rpm (maximum) for 30 min. A 60-min destaining period in running
buffer (or distilled water) is required if ethidium bromide staining is
used. The multigel is then photographed under ultraviolet light using
standard positive/negative film or using a digital camera, with a UV-
compatible ruler placed next to the edge of the multigel for measurement
purposes.
[2] biophysical analysis of genomic loci 25
Southern Hybridization
To detect specific genomic chromatin bands by Southern blotting (see
later), the gel was subsequently treated with standard denaturing and neu-
tralizing solutions, and transferred to Hybond N or NX membranes using
20 SSC in combination with the Schleicher and Schuell Turboblotter.
The use of the Turboblotter (transferring from top to bottom) avoids
having to flip the multigel, and in doing so minimizes the risk of separation
of the running gels from the agarose frame. After transfer, a probe covering
the desired genomic region is radioactively labeled by random priming or
nick-translation. The unincorporated radioactivity is removed using gel fil-
tration (BioSpin P30, BioRad) and the recovered probe is subsequentially
phenol-chloroform extracted. The Hybond membrane is pre-hybridized at
42 for 1.5 h in hybridizing solution (6 SSC, 5 Denharts, 0.5% SDS,
50% formamide, and 0.1 mg/ml denatured salmon sperm DNA). The
probe is boiled for 10 min, added to the hybridization solution, and then in-
cubated at 42 overnight. The membranes were washed twice in 2 SSC,
0.1% SDS at 42 and 30 , respectively, and once with 0.2 SSC, 1% SDS
at 27 . The membranes were then exposed to film or to a PhosporImager
screen for 2–5 days.
Data Analysis
For each individual band detected by Southern blotting or fluorescent
staining, the migration in centimeters is measured from the center of the
well to the center of the band using NIH Image17 or equivalent software.
The digitized image of the ruler placed adjacent to the multigel (see
earlier) serves as reference to set the scale (number of pixels per centi-
meters). Migrations (in centimeters) are subsequently converted to (in
cm2/V s). The gel-free mobility, 00 is obtained by extrapolating the linear
region of a plot of log versus agarose percentage to the y-axis. For
an 2.5-kb chromatin fragment, the linear region falls in the range of
0.2–0.9% agarose. The correlation coefficients of the linear regressions
generally are 0.99. Note that the linear region must be determined
empirically for different sized chromatin fragments.
Using Eq. (1), the Pe of each running gel is determined from the meas-
ured and 00 and known Re (30.1 nm) of the bacteriophage T3 internal
standard. The measured Pe was compared to the range that was deter-
mined based on calculated Pe values from 10 random multigels (see
17
R. R. O’Neill, L. G. Mitchell, C. R. Merril, and W. S. Rasband, Appl. Theor. Electrophor. 1,
163 (1989).
26 chromatin proteins [2]
TABLE II
T3
0 and E
4
T3
0 (10 cm2/V s) E (105 cm2/V s)
Table III). If the Pe values fall outside the expected ranges, the data may be
suspect. The Re of the genomic chromatin band(s) in each running gel is de-
termined from their measured and 00 and the calculated Pe for that gel.
To convert the measured 00 to the actual 0, one must correct for the
effects of electroosmosis on the measured (E). This requires empirical
determination of the E for each specific type of commercial agarose,
which is a time-consuming and expensive process. Toward this end, we
have determined the E for the Research Organics low EEO agarose men-
tioned earlier, and found it to be constant over several lots and many years.
By using this specific agarose in each multigel frame and running gel, ex-
perimental measurements of E can be avoided, and the 00 is converted
to the true 0 according to the following equation
0 0 T3
0 ¼ ð0 ÞðT30 þ E Þ=0 E (2)
where the T3
0 also is a measured constant in the same agarose (see Table II
for T3
0 and E values determined in E and E þ 2 mM Mg2þ buffers).
TABLE III
Empirically Determined Pe Range (nm)
0.2 400–600
0.3 300–460
0.4 200–350
0.5 150–300
0.6 150–250
0.7 150–200
0.8 130–200
0.9 120–190
1 100–150
1.3 90–120
1.6 75–100
1.9 65–80
2.2 60–70
2.5 50–55
2.8 45–50
3 37–43
Flexibility Assay
A relative assay for the conformational flexibility of a chromatin frag-
ment is provided by the slope of a graph of Re versus Pe in the gel range
where Re approaches Pe, that is, in high percentage agarose gels.5,6,12 A
slope near zero is indicative of a particle whose conformational features
are not deformed during electrophoresis. Such a particle is relatively inflex-
ible, for example, unfolded nucleosomal arrays in low salt.5 In contrast, the
Re of a more flexible particle (e.g., DNA, irregularly spaced nucleosomal
arrays in low salt) decreases as the Re approaches Pe. For a 2.5-kb chroma-
tin fragment, this occurs between 1.2% and 3.0% agarose. Thus, relative
conformational flexibility is assayed in high-concentration multigels.
Specifically, to confirm the identity of the bands and determine the pre-
cise distance of migration, in one gel, the DNA was transferred to Hybond
NX membrane and probed with the fragment-specific probe. The second gel
is stained with Commassie (GelCode Blue Stain, Pierce) according to the
manufacturer’s specifications. Slices with 1.5 mm-thickness containing the
bands of interest are excised, soaked in 1% SDS, 50 mM Tris-HCl (pH
6.8) and 1% -mercaptoethanol for 15 min at room temperature. The slices
are then transferred to a Pyrex tube containing the same buffer pre-heated
to 95 for 25–35 s (or until the agarose slice becomes translucent), and the
tube immediately chilled on ice. The cooled agarose slice was rapidly placed
in the well of 1.5-mm thick 4–20% SDS gel and electrophoresed at 100 V
for 90 min. The proteins were subsequently transferred to Immobilon-P
PVDF (Amersham) membranes and immunoblotted according to standard
Maniatis western blotting conditions using appropriate antibodies.
Data Interpretation
The final aspect of the AME approach is to interpret the combined
biophysical and compositional data. For illustrative purposes, this section
summarizes selected important observations from our work with genomic
MMTV promoters under different states of transcriptional regulation.10
For example, the inactive form of the genomic MMTV promoter in
mouse 3134 cells was found to contain histone H1, and showed the same
behavior in the AME assays as model H5-containing nucleosomal arrays
assembled in vitro from pure components. Exposure of cells to dexametha-
sone leads to transcriptional activation in vivo.18–20 The transcriptionally
active genomic MMTV promoters isolated from dexamethasone-treated
cells contained RNA Pol II, TBP, Octl, Brgl, and acetylated H3 tail
domains, and its Re increased to 43 nm (compared to 26 nm in the tran-
scriptionally inactive state). Thus, the biophysical measurements support
the conclusion that H1 had been replaced with several large transcription
factors and multiprotein assemblages after hormone treatment. Interest-
ingly, both the inactive and active forms of the genomic MMTV promoter
were capable of forming salt-dependent secondary chromatin structure.
Together, these data support a model in which transcriptional acti-
vation is associated with reorganization of secondary chromatin structures,
18
H. Richard-Foy, F. D. Sistare, A. T. Riegel, S. S. Simons, Jr., and G. L. Hager, Mol.
Endocrinol. 1, 659 (1987).
19
H. Richard-Foy and G. L. Hager, EMBO J. 6, 2321 (1987).
20
T. K. Archer, C. J. Fryer, H. L. Lee, E. Zaniewski, T. Liang, and J. S. Mymryk, J. Steroid
Biochem. Mol. Biol. 53, 421 (1995).
[3] visualization and 3D structure determination 29
General Considerations
With any imaging technique, the information retrievable is limited
by the quality of the sample, and for chromatin, the most important
considerations are as follows.
Purity
The proportion of the sample constituting the structure of interest
should be monitored. A higher level of contamination can be tolerated if
the material under study is structurally distinct. For example, a circular
General Considerations
With any imaging technique, the information retrievable is limited
by the quality of the sample, and for chromatin, the most important
considerations are as follows.
Purity
The proportion of the sample constituting the structure of interest
should be monitored. A higher level of contamination can be tolerated if
the material under study is structurally distinct. For example, a circular
TABLE I
Concentration
A chromatin sample containing 50 g/ml of the material of interest will
usually suffice. Concentrations much lower than this will always yield
poorer images, and cryo-imaging will be all but impossible. Small volumes
may be concentrated and the buffer components exchanged if needed using
Minicon or Centricon (Amicon #42409, 4208) systems.
Buffer Components
Materials that alter the surface tension or otherwise interfere with ad-
hesion or staining, such as detergents, sucrose, or glycerol, must be re-
moved by dialysis, and the sample separated from ‘‘carrier’’ proteins such
as BSA. If the sample is to be fixed (see later), Tris-type buffers must be
avoided as they react with aldehyde fixatives. We prefer HEPES or PIPES,
at the lowest concentrations that provide adequate buffering. For chroma-
tin, the concentrations of monovalent and divalent salts have a strong influ-
ence on compaction, and will be dictated by the experiment. Protein
complexes that require high salt (over 200 mM monovalent ions) for sta-
bility are best fixed at the optimum salt, after which dilution or dialysis can
be used to achieve the appropriate ionic strength.
[3] visualization and 3D structure determination 31
Fixation
With the exception of cryo-imaging, all techniques require the adhesion
of the sample to a support, a process that exposes chromatin to denaturing
surface forces. Fixation in 0.1% EM-grade glutaraldehyde for 4 h followed
by its removal by dialysis is optimal for most chromatin samples. This treat-
ment preserves the sedimentation constant of chromatin after a post-
fixation change in salt concentration. Longer fixation may result in an
irreversible increase in compaction. For small volumes, minidialysers
(Pierce #69570) are very convenient for this procedure.
Positive Staining
This method is primarily useful for verifying the quality of decondensed
chromatin samples, allowing the investigator to check nucleosome number
and distribution, as well as the overall configuration of the array (circular
or linear). A simple but reproducible procedure is as follows:
1. On a freshly exposed sheet of Parafilm, place a drop of sample
(5–10 l), 3 drops (100 l) of water, 3 drops of 2% aqueous uranyl
acetate (UA), and a final drop of water.
2. Place a glow-discharged carbon film on the sample.
1
C. L. Woodcock, L. Y. Frado, G. R. Green, and L. Einck, J. Microsc. 121, 211 (1981).
32 chromatin proteins [3]
3. After 5 min, remove the grid, blot from one side with a flag of filter
paper (verifying visually that the surface remains fully wetted) and
place on the first drop of water for 10 s. Remove, blot, and place on
the next drop, and continue at 10 s intervals through the 3 drops of
stain to the fourth drop of water, which serves to remove unbound
stain. After the grid has air-dried, it is ready for EM observation.
This method provides a faint positive staining, primarily of DNA, and
the microscope should be set up to maximize contrast. Alternatively, the
EM may be set up for tilted beam dark field, which produces excellent
reversed contrast.
A positive staining technique that provides significantly more contrast
and stains protein and DNA but, in our hands, is less reproducible, is to
use ethanolic phosphotungstic acid (PTA). This method is described in
detail elsewhere.2
2
C. L. Woodcock and R. A. Horowitz, Methods Cell Biol. 53, 187 (1998).
3a
J. M. Tyler and D. Branton, J. Ultrastruct. Res. 71, 95 (1980).
[3] visualization and 3D structure determination 33
Fig. 1. Examples of the options for EM imaging of chromatin described in the text. In
this case, the methylated DNA-binding protein MECP2 was bound to defined 12-mer
nucleosome arrays (in collaboration with P. Georgel, P. Wade, and J. Hansen) and fixed with
glutaraldehyde. (A) Shadowing after vacuum drying from glycerol preserves the 3D shape of
the complex and permits measurement of height (between white lines) and diameter (between
black lines). Bar ¼ 30 nm. (B) Negative staining with uranyl acetate gives an overall sense of
size and shape, but requires tomographic reconstruction to reveal details of internal
substructure (C, D). A central slice from a reconstruction (C) reveals the outline of individual
nucleosomes (arrow). Slice-by-slice examination of the reconstructed volume allows the
location and orientation of individual nucleosomes in the array to be identified, and
construction of a solid model (D) of the array, in which nucleosomes are represented by 5
10 nm disks. (E) ECM imaging of unstained particles yields low-contrast images of the
solution conformation, without flattening or drying. Individual nucleosomes can be identified
(arrow). Stereo-pair reconstruction3b of one particle (F) yields the 3D coordinates of
individual nucleosomes (arrow), and a solid model (G) can be built from those coordinates. A
stereo pair does not give orientation information, so nucleosomes are represented as 10 nm
diameter spheres. Bar ¼ 10 nm.
3b
A. Beorchia, M. Ploton, M. Menager, M. Thiry, and N. Bonnet, J. Microsc. 163, 221 (1991).
34 chromatin proteins [3]
4
C. L. Woodcock, H. Woodcock, and R. A. Horowitz, J. Cell Sci. 99, 99 (1991).
5
C. L. Woodcock and W. Baumeister, Eur. J. Cell Biol. 51, 45 (1990).
[3] visualization and 3D structure determination 35
6
J. Bednar and C. L. Woodcock, Meth. Enz. 304, 191 (1999).
7
C. L. Woodcock and R. A. Horowitz, Methods 12, 84 (1997).
8
J. Hesse, H. Hebert, and P. J. Koeck, Microsc. Res. Tech. 49, 292 (2000).
[3] visualization and 3D structure determination 37
in negative stain, this is in the range of 10–15 e/A2 per exposure. Many
samples can tolerate up to 25 e/A2. Some proteins and DNA-protein com-
plexes, dependent on a poorly understood combination of the preparative
method and the negative stain, are extremely sensitive to damage. The mi-
gration of stain molecules, producing granulation, is usually obvious to the
eye, but damage to the biological material is subtler. If the microscope is
equipped with a CCD camera for digital data recording, it is useful to
collect a series of extremely low-dose images of a single field, until the cu-
mulative dose is 25 e/A2. The images can then be analyzed by calculating
difference images (arithmetic pixel-by-pixel subtraction of one exposure
from the next) to determine the dose at which damage becomes detectable.
Other factors that must be considered are the target resolution in the
reconstruction, the size and distribution of the protein on the support film,
and the area of the CCD chip. Film collection permits hundreds or thou-
sands of particles to be selected from a single exposure, whereas
the number of particles in a single CCD image is far fewer. However, film
collection necessitates scanning the negatives into digitally accessible data
at the appropriate resolution. The pixel size should be 3–4 smaller than
the desired resolution. It is possible with either scanners or CCD imaging
to greatly oversample; this should be avoided as artifacts can be introduced
when data management requires interpolation to larger pixel sizes. In
negative stain, the final resolution in a reconstruction of an asymmetric
particle is unlikely be better than 25A, necessitating a pixel size in the
raw data of 6–8A. For the yeast SWI/SNF complex, this was achieved using
a 2048 2048 CCD camera, each CCD image containing 50–100 particles.
Without convenient access to a high-resolution film scanner, the time
invested in collecting larger numbers of CCD images may be offset by the
time savings of eliminating darkroom work and the digitizing of negatives.
Microscope Parameters
With the target resolution determining the final pixel size, and there-
fore the magnification at which to record the images, other microscope pa-
rameters need be determined. While the best resolution will be achieved at
intermediate voltages in field emission gun (FEG)-equipped microscopes,
perfectly acceptable data can be acquired with care with LaB6 filament–
equipped microscopes at 80–120 kV. The highest resolution will be
obtained at the highest available voltage. The illumination conditions
(small condenser aperture, coordination of spot size with spread of beam)
should be optimized to provide the most coherent beam. The defocus
for the exposures should be selected so that the first zero of the contrast
transfer function (CTF) falls beyond the target resolution.
38 chromatin proteins [3]
Specimen Preparation
Many protein complexes, including ySWI/SNF, are unstable in the con-
ditions required for successful EM and must be chemically cross-linked
(see earlier). For this sample, we used glutaraldehyde fixation following
removal by dialysis of the glycerol and high salt required for long-term
stability, and PTA proved to be the most suitable negative stain.
Data Preparation
The first step in calculating a single-particle reconstruction is to select the
individual particles. First, each source image (negative or CCD image) must
be evaluated for quality: images with drift or astigmatism are rejected and
[3] visualization and 3D structure determination 39
groups of images with like defocus/CTF grouped together. Then the individ-
ual particles are excised from the parent image and centered in individually
‘‘boxed’’ fields. In order for subsequent programs to reorient the selected
particles, the box size should be 25–50% larger than the particle itself.
Most of the image processing packages that include SPR: SPIDER
(www.wadsworth.org/spider), EMAN (ncmi.bcm.tmc.edu), IMAGIC
(www.ImageScience.de), and SUPRIM (ami.Scripps.edu), offer the option
of automating the particle selection process. This can be done by cross-cor-
relation with projections of a previously calculated model if one exists. A
semi-automated method uses a subset of user-identified particles, which
are rotationally averaged and then used as a template for cross-correlation
with the particles in the raw image. This was used successfully for ySWI/
SNF; with sufficiently relaxed criteria, the program will ‘‘overpick’’ and
then the operator needs only to delete improper selections, a considerably
shorter task than hand-selecting each particle.
9
J. Ruprecht and J. Nield, Prog. Biophys. Mol. Biol. 75, 121 (2001).
10
P. A. Thuman-Commike, FEBS Lett. 505, 199 (2001).
11
C. L. Smith, R. A. Horowitz-Scherer, J. F. Flanagan, C. L. Woodcock, and C. L. Peterson,
Nat. Struct. Biol. 10, 141 (2003).
12
S. J. Ludtke, P. R. Baldwin, and W. Chiu, J. Struct. Biol. 128, 82 (1999).
13
M. van Heel, Ultramicroscopy 21, 111 (1987).
40 chromatin proteins [3]
reprojected, and the members of the initial classes are then compared to
this set of projections and reassigned to new classes based on their fit to
the projections. The new classes are averaged, and an iterative process of
model construction, reprojection, and reclassification is begun.
Fig. 2. Class averages: angular distribution and comparison to projections. (A) The dots
in the triangle represent a well-spread distribution of orientations the particles have been
classed into. The apex of the triangle is the view down the z-axis (top) of the particle, and the
base represents views along the x-axis (edge). The brighter the dot, the more particles have
been assigned to that orientation. If the dots appear clustered with gaps in the distribution, the
particle has probably assumed a preferred orientation on the EM support film. (B) A good
indicator of convergence is obtained by noting a positive match when comparing the class
averages (top row) with projections of the refined model at the same orientation (bottom
row). If an arithmetic subtraction is performed between the two, the resulting difference
image should consist of undifferentiated noise.
[3] visualization and 3D structure determination 41
iterations are minimized, and are monitored by examining the Fourier shell
correlation (FSC) between rounds (Fig. 3A and B).
The resolution of the reconstruction can be estimated by dividing the
data into halves, for instance, by separating then averaging all the even
and odd numbered images in all the classes then independently recon-
structing from the even/odd averages. A FSC between the two reconstruc-
tions is examined (Fig. 3C), and the resolution at the correlation value of
0.5 provides a reasonable estimate of the resolution of the reconstruction.14
If the number of images comprising each class is very small, indicating
either sample or staining heterogeneity, the half-averages will be very
noisy, and the resolution may be underestimated.
Visualization
The 3D model can be studied by slicing through the volume and exam-
ining contour maps of the densities, and by creating isocontour surface
maps to view the outer shape of the molecule (Fig. 4). Selecting the thresh-
old at which to contour a surface is done by most software packages
from the molecular weight of the sample (assuming a protein density of
1.35 g/cc). The ‘‘true’’ surface may be up to 30% of this value. In negative
stain, it is quite difficult to establish the correct cutoff. For instance, in UA
stained objects, there is always the possibility of partial positive staining of
the complex, especially when nucleic acids are present. These will then
have an opposite density from stain-excluding regions and may be misinter-
preted as ‘‘holes.’’ Surfaces with deep pockets that accumulate stain may
also be difficult to delineate accurately. There are a variety of software
packages designed to perform visualization tasks. Chimera (www.cgl.ucs-
f.edu/chimera) and Vis5D (www.ssec.wisc.edu/billh/vis5d.html) are free-
ware packages we have used; AVS (www.avs.com) and Amira
14
I. I. Serysheva, M. Schatz, M. van Heel, W. Chiu, and S. L. Hamilton, Biophys. J. 77,
1936 (1999).
42 chromatin proteins [3]
15
J. F. Conway and A. C. Steven, J. Struct. Biol. 128, 106 (1999).
[3] visualization and 3D structure determination 43
17
P. Penczek, M. Marko, K. Buttle, and J. Frank, Ultramicroscopy 60, 393 (1995).
18
M. Radermacher, in ‘‘Electron Tomography’’ (J. Frank, ed.), p. 91. Plenum, New York,
1992.
[3] visualization and 3D structure determination 45
Data Collection
Most specimen holders can only be tilted through 60 . High-tilt
holders for tomography, which range from 70 to 80 are available for
room temperature or ECM work (Gatan, www.gatan.com, Fischione,
www.fischione.com). There are also holders designed for double-tilt data
collection, in which after the initial tilt series is collected, the sample is
rotated 90 in the holder and a second series collected.
For acquiring tilt series of negatively stained or frozen-hydrated
specimens, the ability to record extremely low-dose images is required.
For this, a high-sensitivity CCD is invaluable, as images with electron doses
of <1 e/A2 can be recorded. The dose required for a single image with an
optimal damage should be distributed signal/noise ratio (S/N) without spe-
cimen among all the images of a tilt series, resulting in an equivalent S/N in
the back-projected reconstruction.19 With this approach, it is possible to
obtain high-quality reconstructions from images in which the sample is
barely discernible.20
The number of tilt images needed is determined by the desired reso-
lution in the reconstruction, but may be constrained by the beam sensitivity
of the sample. A guideline for the number of tilts required is given by
n ¼ t/r, where t is the thickness of the sample, r is the desired resolution,
and n is the number of equally spaced tilts required.21 A series of low-dose
images taken sequentially over the same field can supply information on
the maximum dose the sample can withstand. The low S/N of images
recorded at 1 e/A2 makes it highly desirable to use automated data
collection, where computer control of the microscope enables successive
fields to be accurately positioned via cross-correlation, and auto-focusing
is available.22
We employ a strategy for data collection which permits the collection of
at least one full 5 tilt series, then a 2.5 interleaved series, and finally, if de-
sired, a 1.25 series. An alternative to using equidistantly spaced angles in
the tilt series is to collect at a more finely spaced increment in the high
tilts.21 This method may be advantageous for very thin, flat samples to
increase resolution in the z-direction of the reconstruction.23
Recording images:
1. In Search mode, select an area to record, preferably in the center
1/3 of the grid.
19
B. F. McEwen, K. H. Downing, and R. M. Glaeser, Ultramicroscopy 60, 357 (1995).
20
R. A. Horowitz, A. J. Koster, J. Walz, and C. L. Woodcock, J. Struct. Biol. 120, 353 (1997).
21
R. A. Crowther, D. J. DeRosier, and A. Klug, Proc. Royal Soc. London A 317, 319 (1970).
22
A. J. Koster, H. Chen, J. W. Sedat, and D. A. Agard, Ultramicroscopy 46, 207 (1992).
23
W. O. Saxton, W. Baumeister, and M. Hahn, Ultramicroscopy 13, 57 (1984).
46 chromatin proteins [3]
also in use. Software environments such as SPIDER and SUPRIM have ex-
tensive support for tomographic reconstruction, as do dedicated packages
such as IMOD (http://bio3d.colorado.edu/imod). Several national re-
sources (see later section) have significant tomography development pro-
jects underway, and can provide expert consultation, software, and access
to state-of-the-art equipment.
24
C. L. Woodcock, in ‘‘Electron Tomography’’ (J. Frank, ed.), p. 313. Plenum, New York,
1992.
25
R. A. Horowitz, D. A. Agard, J. W. Sedat, and C. L. Woodcock, J. Cell Biol. 125, 1 (1994).
48 chromatin proteins [4]
computationally with the typically large data sets involved. An ideal com-
putational visualization environment would permit access to the greyscale
data while having a transparent volume to interactively move on arbitrary
as well as orthogonal axes, and incorporate 3D measurement tools. While
freeware volume visualization packages are available, currently each
falls short in some aspect. Our laboratory currently uses the commercial
package, Amira (www.tgs.com).
National Resources available in the United States.
1
F. J. Asturias and R. D. Kornberg,J. Biol. Chem. 274, 6813 (1999).
2
E. E. Uzgiris and R. D. Kornberg, Nature 301, 125 (1983).
computationally with the typically large data sets involved. An ideal com-
putational visualization environment would permit access to the greyscale
data while having a transparent volume to interactively move on arbitrary
as well as orthogonal axes, and incorporate 3D measurement tools. While
freeware volume visualization packages are available, currently each
falls short in some aspect. Our laboratory currently uses the commercial
package, Amira (www.tgs.com).
National Resources available in the United States.
1
F. J. Asturias and R. D. Kornberg, J. Biol. Chem. 274, 6813 (1999).
2
E. E. Uzgiris and R. D. Kornberg, Nature 301, 125 (1983).
3
J. Frank, ‘‘Three-Dimensional Electron Microscopy of Macromolecular Assemblies.’’
Academic Press, San Diego (1996).
4
F. J. Asturias, W. H. Chung, R. D. Kornberg, and Y. Lorch, Proc. Natl. Acad. Sci. USA 99,
13477 (2002).
50 chromatin proteins [4]
5
G. W. Tischendorf, H. Zeichhardt, and G. Stoffler, Mol. Gen. Genet. 134, 187 (1974).
6
G. Stoffler and M. Stoffler-Meilicke, in ‘‘Modern Methods in Protein Chemistry’’
(H. Tesche, ed.), p. 409. De Gruyter, Berlin (1983).
[4] electron microscopic analysis 51
will be retained and the specimen will resemble a clean carbon film.
Finally, an excessive amount of material adsorbed to the carbon support
film will result in samples where staining is evident, but no individual par-
ticles can be distinguished when an area is examined at higher (e.g.,
35,000) magnification. After determining that the staining and protein
concentration in the samples are adequate, the next step is to collect
images to evaluate the preservation of the particles and the distribution
of particle orientations in the samples. Since individual particles must be
selected for image analysis, it is important to optimize their distribution
by carefully adjusting the protein concentration used during sample prep-
aration. Ideally, particles should be evenly distributed on the support film,
and the separation between individual particles should correspond to 2–3
particle diameters. This will result in a large number of particles being in-
cluded in recorded images, while ensuring that it will be possible to select
individual particles without interference from neighboring ones.
Most of the data for single particle analysis are still collected on film,
but the use of electronic detectors is increasing, and the trend will undoubt-
edly continue. The arguments that follow apply to either type of data, as
the distinction between them largely disappears after film data has been
digitized. Zero-tilt images of the specimens are necessary to assess particle
preservation and orientation. Images for data analysis are typically
recorded at magnifications in the range 35,000–65,000. In the case of film
data, the images should be scanned using a pixel size equivalent to about
one fourth of the desired resolution. The number of pixels available for
alignment of individual particle images is often the more critical parameter
for determining how images should be digitized.
Fig. 1. Structure of RSC in projection. Images of RSC particles (5880) preserved in uranyl
acetate were computationally aligned. All particles seemed to be in a similar orientation but
differed in conformation and were sorted into homogeneous classes using multivariate
statistical analysis and hierarchical ascendant classification.3 A central area of lower density is
apparent in several of the class averages shown, which include 45% of all particles. Most of
the variation in RSC conformation is related to a domain forming the bottom part of the
structure, which is either missing (35% of particles), or collapsed against the top of the structure
(22% of particles).
Fig. 2. Three-dimensional reconstruction of RSC. (A) RSC consists of four modules that
define a central cavity. Two views of the structure (front and back) are shown. (B) The most
significant variation in RSC conformation was due to the collapse (top) or absence (bottom)
of a module that forms the lower part of the RSC structure. The scale bar corresponds to
approximately 100 Å.
that defines the bottom portion of the central cavity. This domain seems to
partially collapse under the sample preservation conditions used in the
study. However, a majority of the RSC particles (45%) show a clearly de-
fined cavity, and that appears to be the most representative conformation
of the complex. Variability in the position of the lower domain may actu-
ally have functional significance, as it could control access to the central
cavity, which would be more accessible when the bottom domain moves
[4] electron microscopic analysis 55
away from the rest of the structure. The shape and size of the central cavity
in the RSC complex matches closely the shape and size of a nucleosome
core particle. Possible binding of a nucleosome in the cavity was tested
by calculating a reconstruction of the RSC/nucleosome complex.
TABLE I
The Number of Particles in an Arbitrary Area is Listed for RSC (First Column),
Nucleosomes (Second Column), RSC in the RSC/Nucleosome Samples (Third
Column), and Nucleosomes in the RSC/Nucleosome Samples. The Last Row
Shows the Average of the Values for Each Column a
RSC/nucleosomes RSC/nucleosomes
RSC Nucleosomes (RSC count) (nucleosome count)
72 70 35 18
65 55 43 21
67 73 41 17
67 57 28 19
72 71 35 27
69 65 36 20
a
Analysis of these numbers indicates that there is an overall decrease in the number of
particles per unit area after mixing of RSC and nucleosomes, probably as a result of
mutual charge neutralization that decreases the affinity of the particles for the
amorphous carbon film support. It also reveals a 40–50% decrease in the expected
number of nucleosomes in the RSC/nucleosome particles, suggesting that about half of
the RSC particles have a bound nucleosome.
with the surrounding density that defines it) corresponding to the tilted par-
ticle images used to calculate the RSC/nucleosome reconstruction was
carried out to attempt to separate empty and occupied RSC particles.
The results from such analysis (not shown) were inconclusive, perhaps be-
cause the area corresponding to the central cavity includes only a relatively
small number of pixels. Incomplete formation of the RSC/nucleosome
complex detected by the particle count study (Table I), along with staining
artifacts produced by the highly charged nature of the histone core and
DNA surfaces, must contribute to the detection of an artificially low
density in the central RSC cavity.
The structure of the RSC/nucleosome complex strongly suggests that
interaction of a nucleosome with RSC occurs by binding of the nucleosome
to the central cavity in the RSC complex. The size and shape of the central
cavity match closely the size and shape of a nucleosome core particle. This
is illustrated in Fig. 5, which shows our proposed model for RSC/nucleo-
some interaction. The RSC complex is shown in its prevalent confor-
mation, and a low-resolution (25 Å) structure of the nucleosome core
particle (obtained by low-pass filtering of the atomic-resolution X-ray
structure of the complex9) has been placed in the central RSC cavity.
9
K. Luger, A. W. Mader, R. K. Richmond, D. F. Sargent, and T. J. Richmond, Nature 389,
251 (1997).
[4] electron microscopic analysis 59
The figure illustrates the close fit of the nucleosome in the cavity. The
model shown in Fig. 5 is supported by results from nuclease digestion stud-
ies of nucleosomal DNA in an RSC nucleosome complex, which indicate
that digestion of nuclesomal DNA occurs preferentially at positions that
would be more exposed in the model.4
10
A. Miyazawa, Y. Fujiyoshi, M. Stowell, and N. Unwin, J. Mol. Biol. 288, 765 (1999).
11
G. Ren, A. Cheng, V. Reddy, P. Melnyk, and A. K. Mitra, J. Mol. Biol. 301, 369 (2000).
12
B. Bottcher, S. A. Wynne, and R. A. Crowther, Nature 386, 88 (1997).
13
J. Dubochet, M. Adrian, J. J. Chang, J. C. Homo, J. Lepault, A. W. McDowall, and
P. Schultz, Q. Rev. Biophys. 21, 129 (1988).
14
J. Frank, P. Penczek, R. K. Agrawal, R. A. Grassucci, and A. B. Heagle, Methods Enzymol.
317, 276 (2000).
15
P. Penczek, M. Radermacher, and J. Frank, Ultramicroscopy 40, 33 (1992).
16
P. Penczek, R. A. Grassucci, and J. Frank, Ultramicroscopy 53, 251 (1994).
[4] electron microscopic analysis 61
of the RSC complex from the unstained particle images entails combining
the information present in this set of projections. This requires precise
knowledge of their relative orientations. Evidently, approximate informa-
tion is provided by comparison with the available stain reconstruction.
However, preliminary analysis indicates that the deformation induced by
preservation in stain is significant, and this prevents direct use of the stain
reconstruction as a meaningful reference for calculation of an improved re-
construction from the unstained particle images. We are currently working
to obtain a suitable reference volume, which will allow us to make full use
of the information contained in images of unstained RSC particles.
Introduction
Over the past decade, optical trapping techniques have become a stan-
dard part of the repertoire of tools available for the study of biological
molecules.1–5 More recently, optical trapping techniques have been applied
to the study of chromatin structure and even details of the structure of
individual nucleosomes in a chromatin array.6–8
The general experimental design of optical trapping experiments with
chromatin involves immobilization of one end of a linear DNA molecule
on a surface, while the other end of the molecule is attached to a polystyrene
microsphere (bead). The microsphere can then be used as a microscopic
‘‘handle’’ which can be captured and manipulated by the optical trap
(Fig. 1). The optical trap can be used to exert and measure piconewton-scale
1
K. Svoboda, C. F. Schmidt, B. J. Schnapp, and S. M. Block, Nature 21, 365 (1993).
2
H. Yin, M. D. Wang, K. Svoboda, R. Landick, J. Gelles, and S. M. Block, Science 270,
1653 (1995).
3
M. D. Wang, M. J. Schnitzer, H. Yin, R. Landick, J. Gelles, and S. M. Block, Science 282,
902 (1998).
4
J. Liphardt, B. Onoa, S. B. Smith, I. Tinoco, and C. Bustamante, Science 292, 733 (2001).
5
S. J. Koch, A. Shundrovsky, B. C. Jantzen, and M. D. Wang, Biophys. J. 83, 1098 (2002).
6
Y. Cui and C. Bustamante, Proc. Natl. Acad. Sci. USA 97, 127 (2000).
7
M. L. Bennink, S. H. Leuba, G. H. Leno, J. Zlatanova, B. G. de Grooth, and J. Greve, Nat.
Struct. Biol. 8, 606 (2001).
8
B. Brower-Toland, R. C. Yeh, C. Smith, C. L. Peterson, J. T. Lis, and M. D. Wang, Proc.
Natl. Acad. Sci. USA 99, 1960 (2002).
of the RSC complex from the unstained particle images entails combining
the information present in this set of projections. This requires precise
knowledge of their relative orientations. Evidently, approximate informa-
tion is provided by comparison with the available stain reconstruction.
However, preliminary analysis indicates that the deformation induced by
preservation in stain is significant, and this prevents direct use of the stain
reconstruction as a meaningful reference for calculation of an improved re-
construction from the unstained particle images. We are currently working
to obtain a suitable reference volume, which will allow us to make full use
of the information contained in images of unstained RSC particles.
Introduction
Over the past decade, optical trapping techniques have become a stan-
dard part of the repertoire of tools available for the study of biological
molecules.1–5 More recently, optical trapping techniques have been applied
to the study of chromatin structure and even details of the structure of
individual nucleosomes in a chromatin array.6–8
The general experimental design of optical trapping experiments with
chromatin involves immobilization of one end of a linear DNA molecule
on a surface, while the other end of the molecule is attached to a polystyrene
microsphere (bead). The microsphere can then be used as a microscopic
‘‘handle’’ which can be captured and manipulated by the optical trap
(Fig. 1). The optical trap can be used to exert and measure piconewton-scale
1
K. Svoboda, C. F. Schmidt, B. J. Schnapp, and S. M. Block, Nature 21, 365 (1993).
2
H. Yin, M. D. Wang, K. Svoboda, R. Landick, J. Gelles, and S. M. Block, Science 270,
1653 (1995).
3
M. D. Wang, M. J. Schnitzer, H. Yin, R. Landick, J. Gelles, and S. M. Block, Science 282,
902 (1998).
4
J. Liphardt, B. Onoa, S. B. Smith, I. Tinoco, and C. Bustamante, Science 292, 733 (2001).
5
S. J. Koch, A. Shundrovsky, B. C. Jantzen, and M. D. Wang, Biophys. J. 83, 1098 (2002).
6
Y. Cui and C. Bustamante, Proc. Natl. Acad. Sci. USA 97, 127 (2000).
7
M. L. Bennink, S. H. Leuba, G. H. Leno, J. Zlatanova, B. G. de Grooth, and J. Greve, Nat.
Struct. Biol. 8, 606 (2001).
8
B. Brower-Toland, R. C. Yeh, C. Smith, C. L. Peterson, J. T. Lis, and M. D. Wang, Proc.
Natl. Acad. Sci. USA 99, 1960 (2002).
Fig. 1. Instrument and experimental configuration (not to scale). (A) Overall design of
the instrument. The drawing is modified from Wang et al.9 (B) Experimental configuration.
Under feedback control, a nucleosomal array is stretched between the surface of a microscope
coverslip and an optically trapped microsphere.
64 chromatin proteins [5]
Preparation of Histones
Owing to the sensitivity of single molecule studies of chromatin, the
use of highly purified biochemical components is critical to the success of
these experiments. Large quantities of highly purified histone proteins
can be prepared from various chromatin samples using standard hydroxya-
patite (HAP) chromatography. Starting with washed nuclei from the tissue
source of choice, their chromatin content is fragmented by mild MNase
digest, and bound to HAP (BioGel HTP, BioRad Laboratories, Hercules,
CA) by virtue of their nucleic acid component as previously detailed.10,11
Linker histone and nonhistone proteins are removed by washing the
HAP bed at moderate ionic strengths. Finally, core histone proteins are
eluted from the HAP:DNA complex at high ionic strength.
Alternatively, highly purified recombinant histone proteins can be
obtained by expression in bacteria, permitting choice and manipulation
of primary sequence.12 This flexibility permits the design of experiments
9
M. D. Wang, H. Yin, R. Landick, J. Gelles, and S. M. Block, Biophys. J. 72, 1335 (1997).
10
A. P. Wolffe and K. Ura, Methods 12, 10 (1997).
11
G. Schnitzler, in ‘‘Current Protocols in Molecular Biology’’ (F. A. Ausubel et al., eds.), 3,
p. 21.5.7. Wiley, New York, 2001.
12
K. Luger, T. J. Rechsteiner, and T. J. Richmond, Methods Enzymol. 304, 3 (1999).
[5] optical trapping techniques 65
13
A. Flaus, K. Luger, S. Tan, and T. J. Richmond, Proc. Natl. Acad. Sci. USA 93, 1370 (1996).
14
B. Brower-Toland and M. D. Wang, unpublished data.
66 chromatin proteins [5]
Chromatin Assembly
Assembly of chromatin arrays for analysis from highly purified DNA
and histone components can be achieved by chemical (gradient salt
dialysis)15 or by enzymatic means.16 The salt dialysis method of assembly
has the advantage of preserving sample purity and minimizing the amount
of post-assembly purification required before experimentation.
Enzymatic assembly of chromatin has the pitfall of introducing a large
number of protein impurities if an assembly extract is utilized. The recent
development by the Kadonaga lab of a completely recombinant assembly
system minimizes this complication.17 Chromatin assembly by this method
is more rapid than by salt dialysis. Moreover, it has the advantage of pro-
ducing extremely regular arrays of nucleosomes in a sequence-independent
fashion, without introducing artifactual structures such as dinucleosomes.
With either technique, optimization of assembly conditions by post-
assembly electrophoretic analysis is necessary prior to single-molecule
experimentation, both to avoid artifactual structures and to ascertain the
quality of array formation.
15
K.-M. Lee and G. Narlikar, in ‘‘Current Protocols in Molecular Biology’’ (F. A. Ausubel
et al., eds.), 3, p. 21.6.3. Wiley, New York, 2001.
16
M. Bulger and J. T. Kadonaga, Methods Mol. Genet. 5, 241 (1994).
17
M. E. Levenstein and J. T. Kadonaga, J. Biol. Chem. 277, 8749 (2002).
[5] optical trapping techniques 67
Instrumentation
Mechanical measurements of a single nucleosomal array can be
obtained by using a single-beam optical trapping microscope.5,8 Here, we
provide a brief overview of the instrument. The reader should refer to
Wang et al.9 and Koch et al.5 for more detailed descriptions of the design,
construction, and calibration of the optical trapping setup.
Data Analysis
Nucleosomes can be disrupted in various ways. The two ways presented
here are velocity-clamp stretching and force-clamp stretching. These
two methods are roughly equivalent, but with some subtle differences. Vel-
ocity clamp allows disruption of all nucleosomes at a specified stretching
velocity regardless of the strength of protein-DNA interactions within
nucleosome. However, nucleosomes in an array are disrupted under
slightly different force conditions, which depend on the number of nucleo-
somes remaining in the array at a specific disruption. Force clamp allows
disruption of all nucleosomes under identical force conditions (i.e., the
same force). However, more experimentation is required to determine a
workable range of force: If the force is too small or too large, the time to
disrupt the nucleosomes will be too long or too short to be experimentally
accessible.
18
J. F. Marko and E. D. Siggia, Macromolecules 28, 8759 (1995).
70 chromatin proteins [5]
Velocity Clamp
An example of data taken with a velocity clamp is shown in Fig. 2. At
low force (<15 pN), the force-extension curve starts to deviate from that of
the corresponding naked DNA; at higher force (>15 pN), a distinctive saw-
tooth pattern starts to appear; at even higher force (>40 pN), the force-
extension resembles that of a naked DNA (dotted curve). Previously,
Brower-Toland et al.8 demonstrated that the high-force sawtooth pattern
is indicative of a nucleosomal array, with each peak corresponding to a
single nucleosome. Under the conditions used in our experiments, the
spacing between adjacent peaks is 27 nm. The observed sawtooth pattern
suggests separate disruption of strong DNA-histone interactions in
individual nucleosomes.
To determine the amount of DNA released from a nucleosome, the
MMS model can be applied. This conversion attributes extension only to
naked DNA, that is, linker DNA and DNA peeled from nucleosome core
particles (NCP). This method of conversion from force-extension curve to
number of base pairs of naked DNA is similar to that previously used for
single-molecule studies of transcription.3 The MMS model is only an
approximation for a nucleosomal array. To achieve better precision for
the conversion, a more refined model will be necessary. Conversion of
the data in Fig. 2A is shown in Fig. 2B, where the amount of naked DNA
is plotted as a function of time during stretching. At the beginning of
stretching (0–2 s), the average amount of naked DNA is constant, indicat-
ing no DNA release from NCPs. This should correspond to the amount
of linker DNA for a relaxed nucleosomal array. As force rises in the
low-force region, DNA release is gradual, indicating a simultaneous release
of DNA from all nucleosomes, with 76 bp of DNA release per nucleo-
some. At high force, the sawtooth peaks in Fig. 2A, are converted to steps.
DNA release is sudden, indicating a separate release of DNA from each
nucleosome of 80 bp.
Force Clamp
An example of data taken with a force clamp at 20.2 pN is shown in
Fig. 3. Unlike the velocity clamp measurements, all the nucleosomes
experienced the same force before disruption. Here, sudden disruptions
of nucleosomes resulted in stepwise increases in the DNA extension, with
full-length naked DNA (dotted line) are shown for comparison. (B) Amount of naked DNA
as a function of time derived from data shown in Fig. 2A. The top dotted line is a comparison
with a full-length naked DNA. At higher force, the curves show 17 steps, which correspond to
the 17 disruption peaks in Fig. 2A.
72 chromatin proteins [5]
Fig. 3. Stretching a nucleosomal array with a force clamp. The graphs are plots of DNA
extension (left axis) and the amount of naked DNA (right axis) versus time under constant
force of 20.2 pN.
Conclusion
Renewed interest in chromatin as a mediator of the structure, mainten-
ance, and regulation of eukaryotic genomes has inspired the development
of a variety of novel chromatin techniques. Optical trapping technology
provides a useful addition to this repertoire of techniques. We anticipate
that single-molecule optical trapping experiments on chromatin structure
will complement more traditional technologies, and aid in the elucidation
of the structural role in chromatin of histone and nonhistone proteins
and their post-translational modifications. Likewise, optical trapping
methods will be adaptable to the study of enzymatic activities such as
RNA polymerases and ATP-dependent chromatin remodelers operating
on chromatin structure.
Acknowledgments
We thank David A. Wacker and Dr. Robert M. Fulbright for their helpful discussions and
technical assistance. Supported by grants from the NIH and the Keck Foundation’s
Distinguished Young Scholar Award to M.D.W.
[6] single-molecule analysis of chromatin 73
1
S. H. Leuba and J. Zlatanova, eds., ‘‘Biology at the Single-Molecule Level,’’ Pergamon,
Amsterdam, 2001.
2
G. Binnig, C. F. Quate, and C. Rohrer, Phys. Rev. Lett. 56, 930 (1986).
tool: in this application, the AFM tip is allowed to move only in the
z-direction, up and down. If a molecule that is immobilized on the surface
is also attached to the tip (either through specific linkages or through
nonspecific adhesion), then the upward movement of the cantilever will
stretch the molecule to produce the so-called force-extension curves.3
Using OT is a technique for manipulating single molecules based on the
interaction of light with matter.4 Light can exert forces on small beads of
certain optical properties in such a way that the bead is kept suspended
at a point close to the waist of a laser beam that is focused with an objec-
tive. If such a ‘‘trapped’’ bead is pulled out of this equilibrium position
by some external force, a net restoring force resulting from the bead’s
interaction with light will effectively pull it back to this restoring position
(Fig. 1B; see ‘‘Physics of Optical Trapping’’ and Fig. 2A–C). Thus, the
trap acts as a Hookean spring, with the force on the trapped bead being
proportional to its displacement from the equilibrium position in the trap
(Fig. 2D). In the OT set-up, one can suspend a macromolecule between
an optically trapped bead and another surface (a coverslip or another
bead), which can be moved at will to apply tension to the molecular tether,
and force-extension curves can be recorded.
Finally, in the MT, the macromolecule is attached between a surface
and a magnetic bead (Fig. 1C). Manipulation of an external magnetic field
can be used to apply stretching force to the tethered molecule, and/or to
induce precisely known levels of supercoiling: stretching is achieved by
changing the distance between the external magnet(s) and the cuvette,
while supercoiling is introduced by rotating the external magnetic field
3
J. Zlatanova, S. M. Lindsay, and S. H. Leuba, Prog. Biophys. Mol. Biol. 74, 37 (2000).
4
A. Ashkin, J. M. Dziedzic, J. E. Bjorkholm, and S. Chu, Optics Lett. 11, 288 (1986).
[6] single-molecule analysis of chromatin 75
Fig. 2. Principle of optical trapping in a single-beam gradient trap. Schematic showing the
propagation of two high-angle rays as they refract at the surface of a high-index spherical
particle, when positioned just below the focal point (A), just above the focal point (B), or just
off-axis within the focal plane (C). The net force resulting from these two rays is always
pointing toward the focal point, establishing a stable trapping point at the focus. Reflection at
the surface of the bead is neglected (if reflection is taken into account, the center of the trap
will be just below the focal point). (D) The force exerted on the trapped bead increases
linearly with the displacement of the bead from the trap center. Conceptually, this is identical
to a Hookean spring with zero rest length. (E) When the trapped bead is attached to a single
DNA molecule, the force exerted by the trap on the bead is equal to the tension within the
molecule. The deflection of the laser beam is used to accurately measure the displacement of
the bead and, therefore, the tension in the molecule. (F) Optical tweezers optical set-up. The
beam of a 2-W continuous-wave infrared laser (1064 nm) is expanded using lens L1
( f ¼ 50 mm) and L2 ( f ¼ 350 mm) (NOTE 9). The diameter of the beam slightly overfills the
back aperture of the water-immersion objective (Obj), which is a necessary condition to
establish a stable optical trap within the flow cell. The intensity of the laser can be tuned by
turning the quarter-wavelength plate (WP) in front of the first lens. A polarizing beam splitter
cube (PBS) just before the objective separates the two orthogonal polarization directions,
enabling fine control of the intensity of the beam at the objective (light that is reflected by the
polarizing beam splitter is stopped using a beam stop, BS). The transmitted (i.e., refracted)
laser light is collected with a 0.9-NA condenser lens (Con) and projected onto a quadrant
detector. Note that the quadrant detector is not positioned in a conjugate image plane
(NOTE 1). This detector quantifies the deflection, which is calibrated and transformed into a
force signal. A neutral density filter (F1) is used in front of the quadrant detector to reduce
the intensity to measurable levels. Within this experimental set-up, the objective is also used
for viewing the beads in the flow cell. The light of a fiber light source is collimated using a 40-
mm lens (L3) and projected onto the back of the condenser, which results in a homogeneous
illumination of the sample. The objective in combination with an additional lens L4
( f ¼ 170 mm) is used to project the microscopic image onto a CCD array. An additional filter
(F2) is used in front of the camera to fully block the infrared laser light. Dichroic mirrors (D1
and D2) are used to separate the laser-beam path from that of the illumination beam.
76 chromatin proteins [6]
AFM imaging
Chromatin fibers are challenging to image by AFM because they are
relatively soft (compared to naked DNA, for example), have features that
are high above the surface, and contain histones that are notoriously sticky.
Softness results in indentation of the sample by the tip, while the sticky his-
tones may easily attach to it: both interfere with proper imaging. Distortion
to the image also occurs when the vertical distance traveled by the tip is too
large—which happens with samples containing topographically high fea-
tures. We have recently published detailed protocols of how to image chro-
matin fibers with the AFM.8 Here, we provide some further thoughts based
on our prior experience with the Digital Instruments (DI, www.di.com)
AFM (e.g., Zlatanova et al. (1994),9 Leuba et al. (1994),10 Yang et al.,11
Zlatanova et al. (1998)12 Leuba, Bustamante, Zlatanova, and van Holde
(1998),13 Leuba, Bustamante, van Holde, and Zlatanova (1998)14), and
our more recent work with the Molecular Imaging (MI, www.molec.com)
instrument (e.g., Karymov et al.,15 Tomschik et al.,16 An et al.17).
5
T. R. Strick, J. F. Allemand, D. Bensimon, A. Bensimon, and V. Croquette, Science 271,
1835 (1996).
6
T. R. Strick, J.-F. Allemand, V. Croquette, and D. Bensimon, J. Stat. Phys. 93, 647 (1998).
7
T. Strick, J. Allemand, V. Croquette, and D. Bensimon, Prog. Biophys. Mol. Biol. 74,
115 (2000).
8
S. H. Leuba and C. Bustamante, in ‘‘Methods in Molecular Biology: Chromatin Protocols’’
(P. B. Becker, ed.), p. 143. Humana Press, Totowa, NJ, 1999.
9
J. Zlatanova, S. H. Leuba, G. Yang, C. Bustamante, and K. van Holde, Proc. Natl. Acad.
Sci. USA 91, 5277 (1994).
10
S. H. Leuba, G. Yang, C. Robert, B. Samori, K. van Holde, J. Zlatanova, and
C. Bustamante, Proc. Natl. Acad. Sci. USA 91, 11621 (1994).
11
G. Yang, S. H. Leuba, C. Bustamante, J. Zlatanova, and K. van Holde, Nat. Struct. Biol. 1,
761 (1994).
12
J. Zlatanova, S. H. Leuba, and K. van Holde, Biophys. J. 74, 2554 (1998).
13
S. H. Leuba, C. Bustamante, J. Zlatanova, and K. van Holde, Biophys. J. 74, 2823 (1998).
14
S. H. Leuba, C. Bustamante, K. van Holde, and J. Zlatanova, Biophys. J. 74, 2830 (1998).
15
M. A. Karymov, M. Tomschik, S. H. Leuba, P. Caiafa, and J. Zlatanova, FASEB J. 15,
2631 (2001).
16
M. Tomschik, M. A. Karymov, J. Zlatanova, and S. H. Leuba, Struct. Fold. Des. 9, 1201 (2001).
17
W. An, V. B. Palhan, M. A. Karymov, S. H. Leuba, and R. G. Roeder, Mol. Cell 9,
811 (2002).
[6] single-molecule analysis of chromatin 77
There are two different modes of imaging with the AFM, contact mode
and dynamic mode. In the contact mode, the AFM tip is in constant contact
with the sample/surface during the raster-scanning. In the dynamic (so-
called tapping) mode, the cantilever is oscillated at a given frequency
above the surface and the effect of the sample is to dampen the amplitude
of this oscillation. In tapping mode, the tip contacts the sample/surface only
intermittently, thus minimizing the friction/drag forces exerted by the tip in
contact mode. The different instruments achieve a controlled oscillation of
the cantilever in different ways (see later).
Various commercial AFM designs have advantages and disadvantages
in the ease and quality of imaging over an area of 0.5 m 0.5 m. The
most useful image size for chromatin fibers is 1 m 1 m, 512 pixels
512 pixels, as it gives sufficient nucleosome resolution to perform measure-
ments of center-to-center internucleosomal distances, angles between lines
connecting three contiguous nucleosomes, z-heights, and numbers of
nucleosomes per 10-nm fiber length.
Digital Instruments has two designs of the instrument: the Bioscope and
the Multimode. In the Bioscope, the piezo and the cantilever holder are
both above the imaging surface. This scope can be mounted on inverted
fluorescence microscopes: one can easily find large samples (e.g., cells or
metaphase chromosomes) on the surface and then position the scanning
tip directly over the region of interest. The Bioscope comes with a piezo
scanner with a range of 100 m 100 m. Since biological processes occur
in structures in the nanometer range, scan sizes of a couple of micrometers
are needed; such needs are better met by the smaller scanners (typically
16 m 16 m, or 12 m 12 m) of the Multimode instrument. Here,
the piezo scanner is below the imaged surface, creating a risk for buffer
leakage over (and destruction of) the piezo. The DI instrument uses
separate piezos to drive the cantilever oscillation in tapping mode; this
adversely affects the instrumental noise.
The MI AFM uses magnetically driven cantilevers to image biological
samples in air and in buffers. An external solenoid drives an alternating
current (AC) signal that vibrates a cantilever coated with a magnetically
susceptible material (the cantilever responds to the changes in the AC
magnetic field). This method of oscillation is much gentler and reduces
the noise in the instrument (for more technical details, see refs. Zlatanova
et al. [2000],3 Lindsay [2001]18). The MI liquid cell is of a simpler design
than that of the DI instrument, and the piezo is placed above the imaged
18
S. M. Lindsay, in ‘‘Scanning Probe Microscopy and Spectroscopy: Theory, Techniques, and
Applications’’ (D. Bonnell, ed.), p. 289. Wiley, New York, 2001.
78 chromatin proteins [6]
19
F. Thoma, T. Koller, and A. Klug, J. Cell Biol. 83, 403 (1979).
80 chromatin proteins [6]
Optical Tweezers
Optics
The instrument is based on an optical microscope equipped with a CCD
camera (Fig. 2F; see also Bennink [2000]20). For the optical trap, a 2-W in-
frared laser is introduced. The 3-mm diameter of the emerging laser beam
20
M. L. Bennink, Ph.D. Thesis, Department of Applied Physics, University of Twente,
Enschede, 2000.
82 chromatin proteins [6]
21
A. Ashkin, Biophys. J. 61, 569 (1992).
22
L. P. Ghislain and W. W. Webb, Opt. Lett. 18, 1678 (1993).
23
L. P. Ghislain, N. A. Switz, and W. W. Webb, Rev. Sci. Instrum. 65, 2762 (1994).
24
S. B. Smith, Y. Cui, and C. Bustamante, Methods Enzymol. 361, 134 (2003).
[6] single-molecule analysis of chromatin 83
and transmitted light (NA ¼ 0.9), and this signal therefore needs experi-
mental calibration of the force.
The position of the quadrant detector has been widely debated within
the optical tweezers community. In our set-up, the detector is positioned
just behind the back focal plane (NOTE 1). The exact position of the
detector behind the back focal plane is not critical, but the beam has to
be within the linear regime of the detector. To determine the linear regime,
you simply put the entire detector onto an x–y translation stage and deter-
mine the signal as a function of the relative position of the beam on the
detector.
The entire set-up is constructed on top of a horizontal vibration-
isolated table to reduce external vibrations (within the building). The beam
pointing stability of the laser is crucial to the measurements (NOTE 2).
Furthermore, the noise due to airflow and dust can be reduced by shielding
the beam and the optical components with a Plexiglas (perspex) box. The
box encloses the beam emerging from the laser all the way to the detector,
leaving an opening at the flow cell to allow easy access. The side branches
for the white light illumination and the CCD array are not critical and
therefore not shielded.
Fig. 3. The flow cell. The custom-made flow cell consists of a sandwich of a standard
microscope glass with two holes (A), two layers of parafilm with a rectangular channel cut out
(B), and a thin cover glass (170 m thick) with the same dimension as the standard
microscope glass. This entire sandwich is positioned onto a metal (i.e., iron) holder (C), and
two plastic brackets are used to clamp the unit onto the holder (D). Inlet and outlet tubing are
glued inside the holes drilled in the plastic brackets. The plastic tubing is pushed over the hole
in the microscope glass and is actually clamping the sandwich on the holder (the plastic
bracket is not touching the glass directly). This also ensures a leak-free connection to the flow
channel. Finally, a second plastic bar (D) is attached on the top of the iron holder, which will
serve as a support on which to glue the inserted micropipette. (E) Exploded view of the flow
cell illustrating how the different parts should be assembled together. M3 nuts and bolts are
used to clamp the brackets holding the flow cell in place and ensure a leak-free connection.
The position of the small capillary between the parafilm and the thin cover slide is also
indicated. The micropipette goes through the capillary into the flow cell, with its tip in the
center of the flow cell. When inserted, the pipette is glued to the horizontal bar, which serves
as a support.
The parafilm is not sticky enough to hold the two glasses together
so you need to heat the sandwich to 60 (NOTE 3) using a clean hot plate.
Put the sandwich flat on the hotplate with the thick glass down. Put a
separate microscope slide on top that allows you to push the entire sand-
wich gently together (if you do not use a second slide, you might break
the thin cover glass). Note the transparency of the parafilm between
the glass slides. It is grey and nontransparent at RT; at 60 it starts
to become transparent and sticky. Push with your finger at the point
where the capillary is integrated until the glass is perfectly flat, that is, no
[6] single-molecule analysis of chromatin 85
distortion in the reflection. Take the sandwich off the plate and let it cool;
the sandwich is sealed now.
To reuse the thick microscope glass with the holes, leave the sandwich
in ethanol for at least a few hours. If you want to use it again right away,
place it in ethanol for 5 min, and put the point of a scalpel between the
two glasses. Keeping the sandwich submerged in ethanol, gently push the
scalpel edge between the two pieces of glass. Give the ethanol time to flow
between the glass and the parafilm, and make sure that the thin glass slide
is not bent too much. The sandwich will fall apart in a few minutes. Use a
cotton-tipped applicator to clean the glass while submerged in ethanol.
For the holder you need:
. Iron plate (2 mm thick), cut as in Fig. 3C
. 4 bolts (M3 10) and nuts (M3)
. 2 bolts (M3 6) and nuts (M3)
. 2 Delrin plastic bars (3 mm thick) with three holes as indicated in
Fig. 3D
. 1 Delrin plastic bar (1.5 mm thick) with two holes as indicated in
Fig. 3D
. 3-mm outer diameter flexible tubing
. Thin fluid line (10–20 cm long)
. Glue (5 min-epoxy)
Place this glass sandwich in the iron holder. The U-shaped holder is cut
out using a laser, but standard metal drilling and cutting tools can be used
as well. Position the sandwich as indicated in Fig. 3E with the thin cover
glass facing the iron. The holes created for the entrance and exit tubing
should be exactly on top of the 7-mm holes within the iron holder. In be-
tween the glass and the iron, you can put one or two layers of parafilm
(not shown in Fig. 3E) to prevent the glass from breaking. Cut out small
holes at the position of the entrance and exit tubing holes in this spacer.
This is important for aligning the tubing later.
Cut the 3-mm exit tubing with a very sharp knife such that the exit
plane is exactly perpendicular to the tube axis; this will ensure a sealed con-
nection to the glass later. The tubing is glued through the center hole in the
bar. For the entrance tubing we use thin tubing for better control of the
flow. This thin tubing is glued within the 3-mm tubing, which is making
the connection to the glass (NOTE 4). This tubing is glued in the center
hole of the second bar.
Use bolts (M3 10) and nuts (M3) to attach the bars over the sandwich
(Fig. 3E). The sandwich is pushed against the holder by the tubing only,
which is enough to keep it in place and to ensure a sealed connection.
When looking from the backside through the 7-mm hole in the holder,
86 chromatin proteins [6]
you can see if the tubing is correctly aligned and sealed. Connect the
second bar (Fig. 3D, top one) on the backside of the holder using two bolts
(M3 6). This will be the holder for the glass micropipette.
Pulling and Placing Micropipettes in the Flow Cell; Mounting the Flow
Cell into the Setup
For micropipettes, we use glass capillaries that we pull using a profes-
sional pipette puller that has four different parameters, which we optimized
to produce long, tapered micropipettes with a tip diameter of 1–2 m (the
diameter of the micropipette is still less than 50 m at 10 mm away from
the tip). Our optimized parameters are: heat, 323; pull, 100; velocity, 10;
time, 240, but they may vary from one puller to the next (even of the same
model), so use these numbers as a starting point and optimize them further.
Start by breaking the glass tube in the middle (wrap the tube in a tissue)
and clamp one piece in the instrument. After pulling, inspect the tip with
a microscope, and if it looks good, gently break off the long tube, again
holding it in a tissue. Then glue the micropipette to a 20- to 30-cm long
3-mm tubing.
View the holder with the glass sandwich, the tubing and the top bar with a
microscope. Use a metal support plate and small magnets to keep the flow
cell in place. The tubing with the micropipette is mounted on a 3D transla-
tion stage, able to move objects with 10-m accuracy over a range of a few
centimeters. You need to move the tip of the micropipette into the square
capillary that is connecting the outside world with the flow cell. Exert ex-
treme care, since if you touch the tip head-on, it will break. Once you have
the tip within the square capillary, you slowly move the pipette in until the
tip is in the middle of the flow cell. Glue the tip to the plastic bar on top using
two-component glue; the flow cell is ready to be mounted in the set-up.
The flow cell is positioned on a Newport 3D translation stage with both
manual and piezo-electric control. An iron plate with a 4-cm diameter
opening, through which the objective accesses the glass surface, is attached
vertically to the stage and the flow cell is attached to it using small cylin-
drically shaped magnets (5 mm in height, 12 mm in diameter, available in
different stores). This allows you to coarsely align the flow cell such that the
tip of the micropipette is in front of the objective (NOTE 5).
Flow System
A semi-automated flow system controls the exchange of buffers within
the flow cell and the flow rate (Fig. 4). The buffers are kept in containers
ranging from 1 ml (for the cell extract) to 50 ml (for the normal flow
buffer). All containers are screw-capped and have two holes drilled into
[6] single-molecule analysis of chromatin 87
Fig. 4. The flow system used to introduce streptavidin-coated beads, biotinylated DNA,
and the cell-free extract at a controlled flow rate. Different containers ranging from 1 to 50 ml
are connected to a closed air-line, whose pressure is controlled using three computer-
controlled pressure valves. These are connected to a high-pressure line (2.0 atm), outside
pressure (1.0 atm) and a barrel, in which low pressure is maintained (0.5 atm). Switching
these valves shortly will change the pressure in the line, and therefore the flow rate in the flow
cell. Thin tubing is used to connect each container to a six-way selection valve, which specifies
which container is to be used. The beads are introduced in the system using a syringe.
the cap. One hole is 3 mm in diameter, and a piece of 3-mm tubing (air-
line) is inserted. A second hole, only 1 mm in diameter, provides access
for a thin fluid line. The connections through the caps are sealed with
5-min epoxy. All airlines are connected to a closed air system. Increasing
the pressure within the airlines will pressurize all the containers and
will maintain buffer flow. The thin tubing is connected to a six-way low-
pressure selection valve (NOTE 6), which enables the operator to manually
select the desired container.
For containers we use transparent disposable vessels:
. 2-ml tube, screw cap, Nalgene cryovials, PP sterile, Nalgene, a
subsidiary of Sybron, Rotherwas, Hereford, UK
. 12-ml tube, blue screw cap, Greiner bio-one, PS tube, sterile,
diameter, 16.8 mm, height, 100 mm, Greiner BV, Alphen a/d Rijn,
The Netherlands
88 chromatin proteins [6]
25
G. H. Leno, Methods Cell Biol. 53, 497 (1998).
[6] single-molecule analysis of chromatin 91
Fig. 5. Schematic presentation of how a single DNA molecule is attached between two
polystyrene beads. Arrows indicate the direction of the flow inside the flow cell. The four lines
forming a cross mark indicate the position of the optical trap. (A) The 2.6-m polystyrene
beads are introduced into the flow cell, and one of them is captured in the optical trap.
(B) The micropipette is moved toward the trapped bead to transfer the bead onto its
tip. Suction is applied to hold the bead in place. (C) A second bead is captured in the optical
trap. (D) Buffer containing DNA molecules replaces the buffer containing the beads. (E) A
single DNA molecule attaches to the trapped bead and stretches out in the flow direction.
(F) The bead on the micropipette is moved at a distance of about 16 m (i.e., length of the
DNA) downstream from the trapped bead to connect to the free end of the DNA molecule.
The presence of the invisible DNA molecule suspended between the two beads is attested to
by moving the pipette away from the trap position and observing the trapped bead being
pulled out of its position.
once a second bead is captured, increase the flow rate to flush the other
beads out of the flow cell. To speed up this procedure, you can position
the micropipette in front of the trapped bead. This will reduce the drag
force on the trapped bead, thus allowing the use of higher flow rates.
When all beads have disappeared, stop the flow completely by turning
the selector valve between two inputs. The signal from the quadrant de-
tector shows a noisy pattern, representing the Brownian motion of the
optically trapped bead. This signal should be centered around zero for both
x and y directions (for this alignment, it is convenient to have the quadrant
detector positioned onto a coarse translation table). Set the flow rate to
1 mm/s (at the position of the trap); a constant force of 30 pN will now
92 chromatin proteins [6]
be detected, which is a direct result of the drag force on the trapped bead.
Switch to the container with DNA molecules, let the DNA solution flow
for 5-10 s, and then reverse to the main buffer. After 1 to 1.5 min, the
DNA is around the position of the trapped bead (Fig. 5D). When a DNA
molecule attaches to the trapped bead, there is a jump in the force because
the molecule causes a significant increase in drag force (10 pN). If no
jump is detected, repeat the procedure until you observe an attachment.
During attachment, a constant flow is maintained to prevent attachment of
the open DNA end to the trapped bead, resulting in a looped molecule.
Only when both ends are attached (see next step) to separate beads can
you stop the flow.
The last step is to attach the second bead (on the micropipette) to the
free end of the DNA molecule. Move the flow cell back and forth such that
the pipette is at a distance of 16 m downstream from the trapped bead
(Fig. 5E). If the molecule becomes tethered, moving the micropipette away
from the trap position will result in a force increase. Be careful not to
destroy the system at this point, since it is rather easy to pull the bead
out of the trap. You can move the micropipette to the other side and pull
the bead out of the trap in an upstream direction (Fig. 5F).
To ensure that just one individual DNA molecule is attached between
the two beads, the force-extension curve is recorded: the force is measured
while the pipette is moved away from the trapped bead. Video analysis
using a frame grabber and custom-written software applying a centroid
method (NOTE 7) is used to determine the distance between the two
beads, that is, the end-to-end distance of a DNA molecule (NOTE 8).
The typical force-extension curve for a single molecule of dsDNA26,27
exhibits a characteristic horizontal force plateau at 65 pN. Having mul-
tiple DNA molecules attached in a parallel fashion would shift the plateau
to higher forces: 130 pN for two molecules, 195 pN for three, etc. Other
features of the force curve, such as the length of the molecule or the force
development at short distances, are used to determine whether the DNA
molecule is free of bound protein molecules. If this curve deviates from
the expected curve, the molecule is discarded.
filter these out resulted in loss of activity of the cell extract). Therefore,
the trap must be turned off (i.e., laser is blocked), and the beads are
kept separate using a constant buffer flow. The tension within the DNA
can be estimated from the flow rate using Stoke’s law, but a correction
has to be made for the shielding effect of the micropipette and the bead
attached to it. An independent method to determine the tension within
the molecule is to determine the Brownian motion of the freely suspended
bead. Video analysis software is used to extract the positions of the
two beads at a 25-Hz time resolution (i.e., video rate). These data are used
to calculate the mean square displacement of the bead perpendicular to
the direction in which the DNA molecule is pulled. This value is directly
related to the force on the DNA molecule, with F ¼ lkBT/<x2>, where
l is the length of the biopolymer, kBT the thermal energy, and <x2>
the mean square displacement.
Thus, the assembly is performed under constant flow conditions and ob-
served as shortening of the apparent end-to-end distance as a result of the
formation of nucleosomal particles28 (240 nucleosomal particles along
the length of the -DNA molecule). Carefully controlling the flow
rate allows the observation of chromatin fiber assembly as a function of
the tension within the DNA.28
To stretch the reconstituted chromatin fiber, the cell extract is replaced
by the main flow buffer at a low flow rate so that the force does not exceed
5 pN. At this point, the trap is turned on, its intensity being slowly in-
creased from zero to ensure that the free bead is gently caught in the trap,
again avoiding large forces on the suspended chromatin fiber. Next, the mi-
cropipette is moved away at a constant velocity (typically 1 m/s) while the
force generated in the fiber is continuously monitored as its length in-
creases. When the distance between the two beads reaches 20–22 m
(the contour length of the ds -DNA molecule is 16 m, so the DNA at
this point is overstretched), the direction of the micropipette movement is
reversed to relax the molecule at the same speed.
The deflection signal representing the force exerted on the chromatin
fiber is measured using a LabView data acquisition card at a frequency of
1.2 kHz; such a frequency allows the observation of very small steps of tens
of nanometers. In the DNA stretching experiments, the end-to-end dis-
tance measurements are performed using real-time video analysis of the
precise positions of the two beads, thus limiting the sampling rate to a max-
imum of 25 Hz. For the chromatin stretching experiments, we needed sam-
pling at higher frequencies. Since the micropipette bead is moved away at a
28
M. L. Bennink, L. H. Pope, S. H. Leuba, B. G. de Grooth, and J. Greve, Single Mol. 2,
91 (2001).
94 chromatin proteins [6]
30
J. F. Marko and E. D. Siggia, Macromolecules 28, 8759 (1995).
31
M. D. Wang, H. Yin, R. Landick, J. Gelles, and S. M. Block, Biophys. J. 72, 1335 (1997).
96 chromatin proteins [6]
distribution function (Fig. 6). This function bins all possible differences
between each pair of calculated contour lengths in the data set and plots
the result as a histogram. The presence of peaks in this histogram is a clear
signature of a quantized step present in the data. Figure 6B reveals a peak
at 65 nm, and multiples thereof (i.e., 130 nm, 195 nm).29 There is also a
strong peak around zero, which is a direct result of the pairwise distribution
function: Many data points that originally are on one rising part of
the curve are reduced to about the same value for the contour length,
producing differences close to zero.
Magnetic Tweezers
Materials
. Square glass cuvettes, 1 mm 1 mm 50 mm (VitroCom,
Mountain Lakes, NJ, USA, www.vitrocom.com)
. 2.8-m streptavidin-coated magnetic beads (Dynabeads, Dynal,
Oslo, Norway, www.dynal.no)
. Steel ball bearing (Allied Industrial Technologies, Cat. No.
100KSFF, H401)
. Planetary-geared 12 V DC motor, 30 rpm (Cramer, Cat. No.
800HN-DC-3277, available from Digi-Key, Thief River Falls, MN,
USA, Cat. No. CRA203-ND, www.digikey.com)
. Elenco Precision regulated power supply, model XP-620 (1.5–15 V)
. Rubber O-rings 6.7 cm 7.3 cm 0.3 cm (2–5/8 in 2–7/8 in
1/8 in) (Allied Industrial Technologies, Cat. No. 01-231 or 01-146)
. NdFeB/NIB magnets (Indigo Instruments, Waterloo, Canada,
www.indigo.com, Cat. No. 33512)
. 0.8-mm inner diameter silicon tubing (BioRad, Hercules, CA, USA,
Cat. No. 731–8210)
. 3-Com HomeConnect video camera (obsolete; try www.ebay.com
alternative digital cameras from Orange Micro, Anaheim, CA,
USA, www.orangemicro.com, and Logitech, Fremont, CA, USA,
www.logitech.com)
. Six-way low-pressure bulkhead selection valve (Upchurch Scien-
tific, Oak Harbor, WA, USA, Cat. No. V-241)
. Low-pressure injection valve (Upchurch, Cat. No. V451)
. Seiwa Optical America objective micrometer (Cat. No. RET-
PCS81X, available from Fisher Scientific, Pittsburgh, PA, USA,
Cat. No. #S11144)
[6] single-molecule analysis of chromatin 97
Instrumental Setup
In the horizontal magnetic tweezers,32 a white light source (either a
fiber optic or a 150-W bulb) is focused onto a square glass cuvette
(Fig. 7A) through a condenser lens, placed roughly halfway between the
light source output and the cuvette (focusing can be facilitated by using a
white postcard to view the beam of light at the cuvette). The beam passes
through a hollow steel bearing on which a rare earth magnet is placed.
The cuvette is held by a machined piece of Delrin plastic bridge (Fig. 7B
and C) bolted onto the three-axis flexure stage. We have fashioned a
paperclip (not shown) to hold the cuvette tightly in place in the Delrin
bridge. The focused light is collected by an objective juxtaposed to the
cuvette. Most experiments were performed with 90 oil (1.2 NA) or 40
air (0.6 NA) objectives. If the objective has a fixed focal plane, typically a
distance of 160 mm, then the video camera is placed 160 mm behind
the objective. Inexpensive ($100) digital cameras can be used. These
cameras typically have either USB or Firewire direct connections to a PC
and thus do not need an extra analog/digital computer board. Public-
domain software such as NIH Image or ImageJ can be used to collect
movies in AVI format.
The magnification of the image on the video screen is determined by
the magnification of the objective and the distance from the camera to
the objective. There are a couple of ways to calibrate the number of nano-
meters per pixel in the image. One can use the dimensions of the 2.8-m
magnetic bead for rough calibration. More accurate calibration can be
done with an objective micrometer; the micrometer is a grid of fine lines
on a glass slide typically separated by distances of 10 m. This gridded slide
is placed at the position of the cuvette, and images are captured with the
video camera located at positions used during an actual experiment. It is
possible to slide the camera toward the objective to increase the field of
32
S. H. Leuba, M. A. Karymov, M. Tomschik, R. Ramjit, P. Smith, and J. Zlatanova, Proc.
Natl. Acad. Sci. USA 100, 495 (2003).
98 chromatin proteins [6]
Fig. 7. Schematic of the horizontal version of the magnetic tweezers set-up and some
technical parts of the magnetic tweezers instrument. This schematic in (A) indicates some
important distances (in millimeters). The hollow bearing with the mounted magnet is movable
up to 50 mm away from the square glass cuvette in order to adjust the force on the magnetic
bead. (B) Side-view drawing of the Delrin bridge to hold the square glass cuvette. (C) Cross-
section at one end of the Delrin bridge. The three 1.4-mm furrows at the top of the piece are
the three possible locations of the cuvette. (D) Side-view drawing of the DC motor-driven
assembly (fabricated out of aluminum plate and Teflon pulleys) for rotating the external
magnet(s) on the embedded bearing. Upper right side of drawing is space on aluminum plate
to mount DC motor that directly rotates the pulley on the right. The two pulleys are
connected by a rubber O-ring. (E) Bird’seye view of the same assembly. In this view it is
possible to see that the pulley on the left is directly connected to the hollow ball bearing that
is mounted on the upper left side of the aluminum plate in (E). Dimensions are in millimeters.
view. This allows observation of a larger number of beads, finding one that
is tethered, centering it on the screen, and then enlarging the image by
manually moving the camera away from the objective. This is a flexibility
advantage of our horizontal MT set-up over the more traditional set-up
with the magnet and the cuvette positioned above the objective of an
inverted fluorescence microscope. In that set-up, the distance between
the eyepiece (or CCD camera) and the objective is fixed (160 mm); thus,
a video image can only be zoomed in and out by changing the objective;
in doing so one often loses the bead of interest.
[6] single-molecule analysis of chromatin 99
Buffers
. Phosphate-buffered saline (PBS) (150 mM NaCl, 3 mM KCl, 2 mM
KH2PO4, 10 mM NaH2PO4, pH 7.2)
. Washing buffer (used to wash out the unattached beads in the
cuvette after it is assembled into the instrument): PBS with 1%
NP40, 0.2% PEG (3350 Da), 1 mM EDTA, 0.2% biotin
33
T. Strick, Ph.D. Thesis, Department de Physique, Ecole Normale Superieure, Paris, 1999.
100 chromatin proteins [6]
. Blue buffer: PBS with blue dextran ( just enough to see the blue),
1% NP40, 0.2% PEG, 2 M NaCl (this buffer is added to the DNA/
beads and used to visualize their entry into the cuvette)
. Histone-dissociation buffer (used to dissociate histones from the
DNA when chromatin fibers are to be disassembled; it can also used
to rinse the DNA/beads in the initial stage of the experiment): PBS
with 0.2% biotin, 1% NP40, 0.2% PEG, 1 mM EDTA, 2 M NaCl
Fig. 8. Schematic to explain how the measurements of travel of the bead across the screen
are done, and how the tether length is calculated. (A) Cross-section of the cuvette with DNA-
tethered magnetic bead and the placement of the external magnet above the incident light;
such a placement tilts the tether so that the angle ’ is about 60 C. (B) Schematic of the
positions of the magnetic bead in the video image on chromatin assembly when the only
applied external force is the magnetic force; the only movement of the bead during assembly
is in the x-dimension of the video screen. We have defined this movement of the bead across
the screen as the ‘‘travel across the screen.’’ (C) Schematic of how the angle ’ is determined
and how travel is converted to actual changes in tether length. The angle ’ is determined by
trigonometry using xinitial and the initial length of the DNA (16 m). The actual tether
length during the assembly/disassembly experiment is then calculated by dividing the travel
(depicted in [B]) by the cosine of ’.
[6] single-molecule analysis of chromatin 103
Notes
NOTE 1: In many set-ups the quadrant detector is positioned in the
back focal plane of the condenser lens, or a position that is optically
conjugate to this. When properly set up, the detector will be completely
104 chromatin proteins [6]
insensitive to lateral displacements of the optical trap and will only sense
displacements of the bead from the trap center. In our set-up, the position
of the trap is fixed, avoiding the need for an optical scheme that will render
the deflection insensitive to trap movements. The quadrant detector is
behind the focal point such that the ring-shaped illumination pattern is
0.5 cm in diameter. This position, however, is not critical.
NOTE 2: We first used a Millennia 10-W CW laser, which has good
beam pointing stability, but is expensive and large. The laser head is
44 cm 15 cm 15 cm and needs to be positioned on top of the optical
table. The controller unit is 60 cm 40 cm 26 cm. Both are connected
with a thick cable, which is not flexible and hard to install. Now, we use a
2-W infrared (1064 nm) laser (CrystaLaser) that is only 16 cm 7 cm
3 cm, with a controller box of only 15 cm 15 cm 5 cm. For a single
trap, a 2-W laser is sufficient to exert forces of 100 pN on a 2.6-m
polystyrene bead.
NOTE 3: To get 60 , start with a hot plate that has not been used for a
while. Put your finger on the plate and turn it on. Put the glass sandwich on
the plate as well. Wait until it feels hot (you have to take your finger off)
and turn off the hot plate. This is about the right temperature you need.
NOTE 4: Cut a piece of 3-mm tubing 1 to 1.5 cm in length. Pull the
thin tubing (10–20 cm in length) through this piece and fill it up with
two-component glue (the thin tubing should be coming through; otherwise
you risk closing the thin tubing with the glue). When solidified, use a sharp
knife to cut the tubing perpendicular to its axis, creating an end that is able
to make a good seal with the glass.
NOTE 5: The high NA objective used for optical trapping and for cre-
ating the microscopic image has a field of view of only 40 m 40 m. This
makes it tedious to find the tip of your micropipette. First align the micro-
pipette (by eye) as well as you can in front of the objective. Then switch to
the microscopic image and use the manual control on the stage to locate
the edge of the flow channel. Scan along the edge until you see the capillary
through which the pipette is entering the flow channel, and follow the
pipette to the tip.
NOTE 6: For connecting tubing to the selector valve, there are small
cone-shaped pieces, provided with the valve. These cones, however, do
not fit the thin tubing that we use. For connecting the thin line, we use
Micro-Lance needles. Using a very sharp scalpel, cut the needle into two
pieces at the point where the thin metal tube ends in the plastic part. This
will provide a smooth end-face in which the metal tubing ends in the center.
Take the plastic gray screws that are on the selector valve and increase the
size of the hole such that the needle is able to go through. The end-face is
still slightly larger in diameter and the needle will stick out of the screw.
[6] single-molecule analysis of chromatin 105
Now attach the thin line to the metal needle and clamp it in the selector by
carefully screwing in the gray unit.
NOTE 7: For accurate position detection of the two beads, custom-
written video analysis routines that use a centroid method have been de-
veloped. Beads imaged by the CCD array appear as light, circular objects
surrounded by a black edge. The bead position is determined as an average
of the positions of all the pixels that are within this black circular edge.
For this, an algorithm starts at a pixel inside this edge and moves slowly
outward in all directions, until a preset threshold level is reached (i.e., that
is lower than the value of the edge). This results in 600–800 pixels that are
next to each other, forming the inner part of the imaged bead. The x- and y-
position of the bead is determined as the average x- and y-position of these
pixels (accuracy of determination of 4 nm).
NOTE 8: The end-to-end distance is obtained by subtracting the radii of
the two beads from the center-to-center distance as determined with the
centroid method (see NOTE 7).
NOTE 9: In a beam-expanding system you normally use two lenses that
are separated at a distance equal to the sum of their focal distances. Within
this optical layout, the distance between the two lenses (L1 and L2 in
Fig. 2F) is only 390 mm, instead of the expected 400. The beam coming
off lens L2 is thus slightly converging. Using such a beam will create a clear
image of the trapped bead, since the bead is trapped stably just below the
focal point. In practice, correct aligning can be done by slightly moving lens
L2 back and forth until the bead is in focus on the CCD array.
NOTE 10: Connecting the tubings in a way that avoids bubble forma-
tion requires practice. Passing bubbles appear to irreparably stick the
tethered beads to the surface, making them useless. It is recommended to
agitate the DNA-connected beads during the deposition process by making
a couple of short flow pulses (turning on and off the flow valve).
Acknowledgments
We thank Drs. Mikhail Karymov, Miroslav Tomschik, Paul Smith, and Guoliang Yang,
and Kirsten van Leijenhorst-Groener, Ravi Ramjit, and Waldemar Koscielny for help with
experiments, Dr. Haocheng Zheng for assistance with figures, and Dr. Jan Bednar for
discussion. This research has been supported by an NCI K22 grant and startup funds from the
University of Pittsburgh School of Medicine (SHL), startup funds from Polytechnic University
(JZ), and the Dutch Foundation of Fundamental Research (FOM) and MESAþ Research
Institute, University of Twente (MLB).
106 chromatin proteins [7]
1
R. Marmorstein, Nat. Rev. Mol. Cell. Biol. 2, 422 (2001).
2
Y. Zhang and D. Reinberg, Genes Develop. 15, 2343 (2001).
3
C. L. Peterson, EMBO Reports 3, 319 (2002).
4
K. E. Neely and J. L. Workman, Biochim. Biophys. Acta 1603, 19 (2002).
5
P. A. Grant, Genome Biol. 2, REVIEWS0003 (2001).
6
B. D. Strahl and C. D. Allis, Nature 403, 41 (2000).
7
S. L. Schreiber and B. E. Bernstein, Cell 111, 771 (2002).
8
J. E. Brownell, J. Zhou, T. Ranalli, R. Kobayashi, D. G. Edmondson, S. Y. Roth, and C. D.
Allis, Cell 84, 843 (1996).
9
R. Marmorstein, Cell. Mol. Life Sci. 58, 693 (2001).
10
D. E. Sterner and S. L. Berger, Microbiol. Mol. Biol. Rev. 64, 435 (2000).
TABLE I
HAT Families and Their Transcription-Related Functions
MYST family H4
Sas2 Yeast Silencing
Sas3 Yeast Silencing (H3)
Esa1 Yeast Cell cycle
progression
MOF Fruit fly Dosage
compensation
Tip60 Human HIV Tat interaction
MOZ Human Leukeomogenesis
HBO1 Human Origin recognition
interaction
TAFII250 family Yeast to TBP-associated H3
human factor
CBP/p300 family Worm to Global All þ non-histone
human coactivator proteins
SRC family Mice and Steroid receptor H3/H4
human coactivators
SRC-1 coactivators
ACTR/AIB1/pCIP/TRAM-1/RAC3
SRC-3
TIF-2
GRIP1
ATF-2 Yeast to Sequence specific H4, H2B
human DNA-binding activator
HAT1 family Yeast to Replication-dependent H4
human chromatin assembly
(cytoplasmic)
a
Only preferred histone substrates are indicated.
11
R. Balasubramanian, M. G. Pray-Grant, W. Selleck, P. A. Grant, and S. Tan, J. Biol. Chem.
277, 7989 (2002).
12
L. A. Boyer, M. R. Langer, K. A. Crowley, S. Tan, J. M. Denu, and C. L. Peterson, Mol.
Cell 10, 935 (2002).
13
P. A. Grant, A. Eberharter, S. John, R. G. Cook, B. M. Turner, and J. L. Workman, J. Biol.
Chem. 274, 5895 (1999).
108 chromatin proteins [7]
14
A. Clements, J. R. Rojas, R. C. Trievel, L. Wang, S. L. Berger, and R. Marmorstein, EMBO
J. 18, 3521 (1999).
15
R. C. Trievel, J. R. Rojas, D. E. Sterner, R. Venkataramani, L. Wang, J. Zhou, C. D. Allis,
S. L. Berger, and R. Marmorstein, Proc. Natl. Acad. Sci. USA 96, 8931 (1999).
16
J. R. Rojas, R. C. Trievel, J. Zhou, Y. Mo, X. Li, S. L. Berger, D. Allis, and R. Marmorstein,
Nature 401, 93 (1999).
17
M. H. Kuo, J. X. Zhou, P. Jambeck, M. E. A. Churchill, and C. D. Allis, Genes Develop. 12,
627 (1998).
18
L. Wang, L. Liu, and S. L. Berger, Genes Dev. 12, 640 (1998).
[7] biochemical characterization of HATs 109
Each of the HAT domains was cloned into the PRSET plasmid vector
for overexpression in Escherichia coli BL21 (DE3) cells. Induction at 37
with IPTG resulted in robust protein induction; however, upon cell disrup-
tion by sonication, the majority of protein was found in the insoluble pro-
tein fraction. In contrast, if the protein was induced with IPTG and induced
overnight at 15 , a significant fraction of the protein (between 25% and
50%) was found in the soluble protein fraction. Although the reason for
this is not clear, it has been proposed that the lower temperature allows
the protein to fold more slowly, resulting in more ‘‘native like’’ folding
and less improper folding that typically leads to protein aggregation.
My laboratory was also able to overexpress the HAT domain from
members of the MYST HAT family that were suitable for biochemical
study19; however, this has not been as straightforward as with the HAT do-
mains from the Gcn5/PCAF family. As with the Gcn5/PCAF HAT
domain, the regions chosen for overexpression relied heavily on sequence
homology among the MYST HAT members (Fig. 1B). Various deletion
constructs of the yeast Esa1 members of the MYST HAT family were pre-
pared and bacterial extracts containing these deletion mutants were
screened for HAT activity. This analysis localized the HAT active domain
to residues 147–445 (the C-terminus of the protein). Induction of a yEsa1
protein construct at 15 , harboring residues 147–445, showed good expres-
sion of soluble protein; however, the protein showed visible degradation
during purification. Inspection of the sequence conservation within the
MYST family revealed that residues 147–162 and 434–445 were poorly con-
served, and based on this we prepared a yEsa1 construct harboring residues
160–435 that was amenable to the overexpression of intact soluble protein
that was suitable for biochemical and structural studies. Subcloning analo-
gous regions of the human homologue of yEsa1, we have more recently
been able to overexpress hMOZ and hTIP60 in bacteria (unpublished).
However, in these cases the proteins were prepared as N-terminal
6 His-fusion proteins and also contained a C-terminal KKK sequence
to improve protein solubility.
19
Y. Yan, N. A. Barlev, R. H. Haley, S. L. Berger, and R. Marmorstein, Mol. Cell 6,
1195 (2000).
110 chromatin proteins [7]
Fig. 1. Sequence homology among HAT proteins. (A) Sequence homology among the
Gcn5/PCAF family of HAT proteins. Residues with black and grey shading are identical and
conserved, respectively. Secondary structural elements within the enzymatic core (black) and
[7] biochemical characterization of HATs 111
N- and C-terminal regions (grey) are color-coded. The ‘‘A’’ and ‘‘H’’ symbols represent
residues that mediate Ac-CoA and histone contacts, respectively, and the catalytic glutamate
general base is indicated with a ‘‘’’ symbol. (B) Sequence homology among the MYST
family of HAT proteins. The color-coding and symbol designation are as described in (A). The
catalytic glutamate and cysteine residues are indicated with a ‘‘’’ symbol.
112 chromatin proteins [7]
20
J. E. Herrera, M. Bergel, X. J. Yang, Y. Nakatani, and M. Bustin, J. Biol. Chem. 272,
27253 (1997).
21
R. C. Trievel, F.-Y. Li, and R. Marmorstein, Anal. Biochem. 287, 319 (2000).
22
Y. Lin, C. M. Fletcher, J. Zhou, C. D. Allis, and G. Wagner, Nature 400, 86 (1999).
23
Y. Yan, S. Harper, D. W. Speicher, and R. Marmorstein, Nat. Struct. Biol. 9, 862 (2002).
24
R. N. Dutnall, S. T. Tafrov, R. Sternglanz, and V. Ramakrishnan, Cell 94, 427 (1998).
[7] biochemical characterization of HATs 113
Fig. 2. Overall structure of HAT proteins. (A) Schematic structure of the Gcn5/PCAF
HAT domain. tGcn5 on complex with CoA (red) and a 19-residue histone H3 peptide (green).
The structurally conserved catalytic core domain is colored in blue, and the structurally
variable N- and C-terminal domains are colored in aqua. (B) Schematic structure of the yEsa1
member of the MYST HAT domain in complex with CoA (red). The color coding is as
indicated in (A). (C) Schematic structure of the yHAT1 HAT domain in complex with Acetyl-
CoA (red). The color coding is as indicated in (A).
Fig. 3. Enzymatic mechanism of HAT proteins. (A) Proposed catalytic mechanism for the
Gcn5/PCAF HATs. Residue numbering for tGcn5 is indicted. (B) Proposed catalytic
mechanism for the MYST HATs. Residue numbering for yEsa1 is indicated.
histone H3 peptide that are ordered in the ternary complex with the 19-residue peptide, but
not ordered in the ternary complex with the 11-residue peptide, are colored in grey. (C)
Superposition of putative substrate binding sites of HAT proteins. The superposition is
generated by superimposing the core domains from the Gcn5/PCAF member, PCAF (pink),
and the MYST member, yEsa1 (blue), with yHAT1 (green). Only the core domain and CoA
of yEsa1 (blue) is shown for clarity.
118 chromatin proteins [7]
residues are ordered with only two disordered residues at each end of the
peptide (Fig. 4A and B) (unpublished). Surprisingly, relative to the com-
plex containing H3p11, the H3p19 complex reveals significantly more ex-
tensive protein-peptide interactions. The greatest structural alterations
between the two peptides occur in residues N-terminal to lysine 14, as
H3p19 makes more interactions in this region. Most strikingly, a core of
12 residues of histone H3 centered on lysine 14 is well ordered in the
tGCN5/CoA/histone H3p19 complex. Each of these residues makes
sequence-specific contacts to protein regions N- and C-terminal to the cata-
lytic core of tGCN5. This comparison reveals that residues outside the core
12-residue sequence of the histone H3 peptide anchor and reposition the
core of histone H3 for more optimal enzyme-histone contacts.
The protein regions N- and C-terminal to the catalytic domains of Esa1
and HAT1 show structural divergence to the corresponding regions of
Gcn5/PCAF (Fig. 2). Despite this, a supposition of the respective core
domains of the two protein families reveals that the N- and C-terminal
regions of Gcn5 that specifically contact histone substrate have structurally
overlapping counterparts in the Esa1 structure (Fig. 4C).19 This suggests
that the Gcn5/PCAF and MYST HATs, and possibly other HAT families,
may have a similar structural scaffold for substrate-specific recognition
and that the sequence divergence within this scaffold may contribute to
substrate-specific binding.
important issue that will require structural studies of relevant HAT com-
plexes. These structures may be at the heart of understanding how HAT
activity is coordinated with other histone-modifying activities to faithfully
modulate transcriptional regulation.
Acknowledgments
I would like to acknowledge all past and present members of my laboratory who have
contributed to the studies discussed in this chapter. In particular, I would like to thank
A. Clements, M. Holbert, A. Poux, J. Rojas, T. Sikorski, R. Trievel, and Y. Yan.
1
T. Pawson and P. Nash, Genes Dev. 14, 1027 (2000).
2
R. D. Kornberg and Y. Lorch, Cell 98, 285 (1999).
3
S. Bjorklund, G. Almouzni, I. Davidson, K. P. Nightingdale, and K. Weiss, Cell 96,
759 (1999).
4
F. Jeanmougin, J. M. Wurtz, B. L. Douarin, P. Chambon, and R. Losson, Trends Biochem.
Sci. 22, 151 (1997).
5
R. Aasland, T. J. Gibson, and A. F. Stewart, Trends Biochem. Sci. 20, 56 (1995).
important issue that will require structural studies of relevant HAT com-
plexes. These structures may be at the heart of understanding how HAT
activity is coordinated with other histone-modifying activities to faithfully
modulate transcriptional regulation.
Acknowledgments
I would like to acknowledge all past and present members of my laboratory who have
contributed to the studies discussed in this chapter. In particular, I would like to thank
A. Clements, M. Holbert, A. Poux, J. Rojas, T. Sikorski, R. Trievel, and Y. Yan.
1
T. Pawson and P. Nash, Genes Dev. 14, 1027 (2000).
2
R. D. Kornberg and Y. Lorch, Cell 98, 285 (1999).
3
S. Bjorklund, G. Almouzni, I. Davidson, K. P. Nightingdale, and K. Weiss, Cell 96,
759 (1999).
4
F. Jeanmougin, J. M. Wurtz, B. L. Douarin, P. Chambon, and R. Losson, Trends Biochem.
Sci. 22, 151 (1997).
5
R. Aasland, T. J. Gibson, and A. F. Stewart, Trends Biochem. Sci. 20, 56 (1995).
6
J. W. Tamkun, R. Deuring, M. P. Scott, M. Kissinger, A. M. Pattatucci, T. C. Kaufman, and
J. A. Kennison, Cell 68, 561 (1992).
7
S. R. Haynes, C. Dollard, F. Winston, S. Beck, J. Trowsdale, and I. B. Dawid, Nucleic Acids
Res. 20, 2603 (1992).
8
J. E. Brownell, J. Zhou, T. Ranalli, R. Kobayashi, D. G. Edmondson, S. Y. Roth, and
C. D. Allis, Cell 84, 843 (1996).
9
P. Filetici, C. Aranda, A. Gonzalez, and P. Ballario, Biochem. Biophys. Res. Commun. 242,
84 (1998).
10
G. A. Marcus, N. Silverman, S. L. Berger, J. Horiuchi, and L. Guarente, EMBO J. 13,
4807 (1994).
11
C. E. Brown, L. Howe, K. Sousa, S. C. Alley, M. J. Carozza, S. Tan, and J. L. Workman,
Science 292, 2333 (2001).
12
D. E. Sterner, P. A. Grant, S. M. Roberts, L. J. Duggan, R. Belotserkovskaya, L. A. Pacella,
F. Winston, J. L. Workman, and S. L. Berger, Mol. Cell. Biol. 19, 86 (1999).
13
T. Georgakopoulos, N. Gounalaki, and G. Thireos, Mol. Gen. Genet. 246, 723 (1995).
14
P. Syntichaki, I. Topalidou, and G. Thireos, Nature 404, 414 (2000).
15
C. Muchardt, B. Bourachot, J. C. Reyes, and M. Yaniv, EMBO J. 17, 223 (1998).
16
C. Muchardt and M. Yaniv, Semin. Cell. Dev. Biol. 10, 189 (1999).
17
P. Chua and G. S. Roeder, Mol. Cell. Biol. 15, 3685 (1995).
18
J. Du, I. Nasir, B. K. Benton, M. P. Kladde, and B. C. Laurent, Genetics 150, 987 (1998).
19
B. R. Cairns, A. Schlichter, H. Erdjument-Bromage, P. Tempst, R. D. Kornberg, and
F. Winston, Mol. Cell 4, 715 (1999).
[8] use of nuclear magnetic resonance spectroscopy 121
Protein Purification
The harvested bacterial cells are resuspended in a lysis buffer [50 mM
Tris-HCl of pH 8.0, containing 10% glycerol, 1% NP-40, 300 mM NaCl,
1 mM PMSF, and EDTA free protease inhibitors (one tablet per liter of
cell culture) (Roche)] and subjected to sonication. Cellular debris is re-
moved by centrifugation at 100,000g for 20 min, and the supernatant
obtained is used for subsequent protein purification.
The bromodomain protein is first purified by affinity chromatography
on a nickel-IDA column (Invitrogen) using an FPLC system (Amersham
Pharmacia Biosciences). The cell lysate is applied to the nickel resin
column of 5 ml (for 3–4 L of cell culture) that is pre-equilibrated in a bind-
ing buffer [50 mM Tris-HCl of pH 8.0, containing 250 mM NaCl, 5 mM
-mercaptoethanol ( -ME), 1 mM PMSF, and protease inhibitors], and
subsequently washed with 10–20 column volumes of the binding buffer,
followed by 10 column volumes of a washing buffer [50 mM Tris-HCl of
pH 8.0, containing 250 mM NaCl, 5 mM -ME, 1 mM PMSF, and protease
inhibitors, plus 20 mM imidazole]. The hexahistidine-tagged bromodomain
protein is eluted from the column in 1.0-ml fractions with an elution buffer
with an 20–500-mM imidazole gradient in 50 mM Tris-HCl of pH 8.0, con-
taining 250 mM NaCl, 5 mM -ME, 1 mM PMSF, and protease inhibitors.
Fractions containing the pure protein are pooled and dialyzed to a throm-
bin cleavage buffer [50 mM Tris-HCl of pH 8.0, containing 250 mM NaCl,
and 5 mM -ME]. The hexahistidine tag in the recombinant protein is re-
moved with thrombin treatment (1 unit thrombin/mg protein) overnight
at 4 . The thrombin cleavage reaction is stopped by addition of 1 mM
PMSF. The protein sample is concentrated and applied to a size exclusion
chromatography column for further purification using an FPLC system in a
phosphate buffer [100 mM phosphate of pH 6.5, containing 0.5 mM EDTA
and 1 mM DTT]. Peak fractions are collected, concentrated to 0.5 mM,
and dialyzed to the final NMR buffer in H2O/2H2O (9:1) consisting of
[8] use of nuclear magnetic resonance spectroscopy 123
25
G. M. Clore and A. M. Gronenborn, Meth. Enzymol. 239, 249 (1994).
26
M. Sattler and S. W. Fesik, Structure 4, 1245 (1996).
27
M.-M. Zhou, K. S. Ravichandran, E. T. Olejniczak, A. P. Petros, R. P. Meadows, M. Sattler,
J. E. Harlan, W. Wade, S. J. Burakoff, and S. W. Fesik, Nature 378, 584 (1995).
28
F. Delaglio, S. Grzesiek, G. W. Vuister, G. Zhu, J. Pfeifer, and A. Bax, J. Biomol. NMR 6,
277 (1995).
29
B. A. Johnson and R. A. Blevins, J. Biomol. NMR 4, 603 (1994).
30
T. Yamazaki, W. Lee, C. H. Arrowsmith, D. R. Mahandiram, and L. E. Kay, J. Am. Chem.
Soc. 116, 11655 (1994).
124 chromatin proteins [8]
31
M. Nilges and S. O’Donoghue, Prog. NMR Spectrosc. 32, 107 (1998).
32
D. Neri, T. Szyperski, G. Otting, H. Senn, and K. Wüthrich, Biochemistry 28, 7510 (1989).
33
G. Vuister and A. Bax, J. Am. Chem. Soc. 115, 7772 (1993).
34
J. C. Matson, O. W. Sörensen, P. Söresen, and F. M. Poulsen, J. Biomol. NMR 3, 239 (1993).
35
G. M. Clore, A. Bax, and A. M. Gronenborn, J. Biomol. NMR 1, 13 (1991).
36
A. T. Brunger, ‘‘X-PLOR Version 3.1: A System for X-Ray Crystallography and NMR.’’
Yale University Press, New Haven, CT, 1993.
37
J. Kuszewski, M. Nilges, and A. T. Brunger, J. Biolmol. NMR 2, 33 (1992).
38
M. Nilges, G. M. Clore, and A. M. Gronenborn, FEBS Lett. 229, 317 (1988).
[8] use of nuclear magnetic resonance spectroscopy 125
hydrogen bonds. Final structure calculations employ the manual and the
ARIA-assisted NOE distance restraints, together with hydrogen bond dis-
tance restraints and dihedral angle restraints. The distance restraint force
constant used in the calculations is typically 50 kcal/mol/Å2, and no NOE
is violated by more than 0.3 Å. The torsion restraint force constant is
200 kcal/mol/rad2, and no dihedral angle restraint is violated by more than
5 . Only the covalent geometry terms, NOE, torsion, and repulsive van der
Waals terms are used in the structure refinement. A large, and negative
Lennard-Jones potential energy should be observed for the final structures,
indicating good non-bonded geometry of the structure. Procheck39 analysis
is also performed to show the majority of the protein residues that are in
preferred and allowed regions of the Ramachandran map. Finally, struc-
tures of proteins determined using NOE-derived distance restraints and
dihedral angle restraints can be further refined with use of residual dipolar
couplings, which can be measured in bicelle-based liquid crystalline or
cross-linked polyacrylamide gel medium and implemented in the final
refinement stage of the structure calculations.40–43
46
R. H. Jacobson, A. G. Ladurner, D. S. King, and R. Tjian, Science 288, 1422 (2000).
47
P. J. Hajduk, R. P. Measdows, and S. W. Fesik, Q. Rev. Biophys. 32, 211 (1999).
48
J. M. Moore, Curr. Opin. Biotech. 10, 54 (1999).
128 chromatin proteins [8]
49
L. Liu, D. M. Scolnick, R. C. Trievel, H. B. Zhang, R. Marmorstein, T. D. Halazonetis, and
S. L. Berger, Mol. Cell Biol. 19, 1202 (1999).
50
W. Gu and R. G. Roeder, Cell 90, 595 (1997).
51
N. A. Barlev, L. Liu, N. H. Chehab, K. Mansfield, K. G. Harris, T. D. Halazonetis, and S. L.
Berger, Mol. Cell 8, 1243 (2001).
52
C. Prives and J. L. Manley, Cell 107, 815 (2001).
53
S. Mujtaba, Y. He, L. Zeng, A. Farooq, J. E. Carlson, M. Ott, E. Verdin, and M.-M. Zhou,
Mol. Cell 9, 575 (2002).
54
M. Sattler, J. Schleucher, and C. Griesinger, Prog. NMR Spectrosc. 34, 93 (1999).
[8] use of nuclear magnetic resonance spectroscopy 129
eluted from the agarose beads are separated by SDS-PAGE and protein/
peptide interaction is visualized by western blotting using anti-GST anti-
body (Sigma) and horseradish-peroxidase–conjugated goat anti-rabbit IgG.
Such a GST pull-down assay is also used to assess other bromodomains’
binding or effect of site-directed mutation of PCAF bromodomain residues
on protein binding to the HIV-1 Tat peptide. Furthermore, this binding
assay can be used for mutational analysis of Tat peptide residues in a
peptide competition assay, in which a non-biotinylated peptide carrying a
mutation at a specific amino acid residue competes with the biotinylated
wild-type Tat AcK50 peptide for binding to the GST-fusion PCAF bromo-
domain. The molar ratio of the mutant and wild-type peptides in the mix-
ture is kept at 1:2. Because of the high sensitivity of western blotting
detection, this binding study can be performed at a protein concentration
(10 M) much lower than that required for NMR study (200 M), thus
ensuring specificity of protein-peptide interactions. Using this GST pull-
down–based assay, mutational analyses of the protein and peptide residues
validate the molecular interactions observed in the NMR structure of
the complex, and identify key residues of the PCAF bromodomain that
are important for recognition of the acetyl-lysine and its flanking residues
in the HIV-1 Tat peptide.
Perspective
The NMR-based methods described here have enabled establishment
of the biochemical functions of bromodomains as acetyl-lysine–binding
domains.20,44,53 This new mechanism of regulating protein-protein inter-
actions via lysine acetylation has broad implications in a wide variety
of cellular processes, including chromatin remodeling and transcriptional
activation.21–23,53,55 Such a powerful NMR structure-based approach is
readily applicable to investigate the biological functions of the other evolu-
tionarily conserved protein modular domains that play an important role in
regulation of chromatin remodeling.
Acknowledgment
The work was supported by a grant from the National Institutes of Health to M.-M. Z.
(CA87658).
55
A. Dorr, V. Kiermer, A. Pedal, H.-R. Rackwitz, P. Henklein, U. Schubert, M.-M. Zhou,
E. Verdin, and M. Ott, EMBO J. 21, 2715 (2002).
[9] interaction of the chromodomain with histone peptides 131
NMR Spectroscopy
In addition to the determination of high-resolution structure, this
technique can be used to rapidly screen for specific protein-protein,
protein-peptide, or protein-small molecule interactions, map the protein
interaction surface, and calculate the binding affinity. All of these applica-
tions have been used successfully to study the interactions of chromodo-
main and bromodomain chromatin binding modules with histone tails.
For example, NMR has been used to determine the specificity of binding of
the HP1 chromodomain to methylated H3 tail peptides and map the surface
of interaction.8 It has also been used to solve the high-resolution structure of
the HP1 chromodomain bound to a methylated histone H3 tail.10
NMR techniques do have limitations. First of all, the protein module
must be able to be expressed in milligram quantities in a few liters of min-
imal media in order to incorporate 15N and/or 13C labeling at an efficient
cost (every liter of minimal media requires 1 g of 15N ammonium chloride
at $30/g and 4 g of 13C glucose at $200/g). In addition, the protein must be
soluble and stable to >200 M in a buffer suitable for NMR measurements
at room temperature (samples of 50–100 M may be studies when using a
state-of-the-art cryoprobe in the NMR spectrometer). Phosphate buffer in
the pH range 6–7.5 is ideal to ensure that backbone amide hydrogens are
fully occupied for maximum sensitivity of detection.
To screen for protein-peptide interactions by NMR spectroscopy, the
main tool is the two-dimensional [1H-15N]-HSQC experiment. The spec-
trum resulting from this experiment correlates the 1H and 15N chemical
shifts, thus providing a backbone fingerprint of the protein. The value of
the chemical shift of each residue is highly dependent on the local chemical
environment of that residue. Thus, addition of a protein or peptide which
specifically binds to the target protein produces a new spectrum with a de-
fined number of large chemical shift perturbations. For example, the ability
of protein modules to bind modified histone tail peptides can be monitored
by titrating an unlabeled peptide into a 15N-labeled protein solution, and
collecting the [1H-15N]-HSQC spectrum of the protein. Figure 1A shows
the [1H-15N]-HSQC spectra of the free chromodomain (black cross peaks)
10
P. R. Nielsen, D. Nietlispach, H. R. Mott, J. Callaghan, A. Bannister, T. Kouzarides, A. G.
Murzin, N. V. Murzina, and E. D. Laue, Nature 416, 103 (2002).
[9] interaction of the chromodomain with histone peptides 133
Fig. 1. NMR analysis of the HP1 chromodomain interaction with the dimethyllysine 9
histone H3 peptide. (A) The [1H-15N]-HSQC spectrum of the free chromodomain (black cross
peaks) and after the addition of the dimethyllysine 9 histone H3 peptide (grey cross peaks).
Each dashed line corresponds to the perturbation of a residue. (B) Chemical shift
perturbation (ave) upon H3 peptide binding for the residues of the chromodomain.
134 chromatin proteins [9]
and after addition of an H3 tail peptide (grey cross peaks) containing a di-
methyllysine at residue 9. The large chemical shift perturbations are indica-
tive of formation of a specific protein-peptide complex. In contrast, an
unmethylated peptide or a peptide methylated at lysine 4 produces a few
small chemical shift changes, indicating that these peptides do not specific-
ally interact with the HP1 chromodomain.8 NMR was also used to map the
surface of the chromodomain which interacts with the H3 tail.8 Figure 1B
shows that a few key residues exhibit the greatest change in chemical shift,
implicating them for direct peptide binding.
X-ray crystallography or NMR spectroscopy of a protein-peptide com-
plex is greatly simplified if only minimally structured domains are studied.
Backbone dynamics determined by NMR can establish which segments of
the protein are ordered into a defined structure. Figure 2 shows a plot of
HP1 chromodomain [1H]-15N NOE values for the free protein (filled tri-
angles) and after complexation with a histone H3 peptide dimethylated
at lysine 9 (open circles). Only residues 25–74 of the chromodomain are
rigid in both the free and bound form as judged by the [1H]-15N NOE value
>0.55. This information was then used to prepare a new compact construct,
which dramatically improved growth of suitable crystals for high-resolution
structure determination by X-ray crystallography.11
11
S. A. Jacobs and S. Khorasanizadeh, Science 295, 2080 (2002).
[9] interaction of the chromodomain with histone peptides 135
Protein Preparation
The coding regions of Drosophila HP1 residues 1–84 and 17–76 were
sub-cloned into NdeI-BamHI sites of pET16b vector (Novagen) and ex-
pressed in BL21(DE3) Escherichia coli (Stratagene) with an N-terminal
His-tag. Isotopic labeling of the protein (15N/13C) was achieved by grow-
ing these cells in minimal media containing 7 g Na2HPO4, 3 g KH2PO4,
0.5 g NaCl, 0.1 g yeast extract, 4 g 13C-glucose, 1 g 15N-NH4Cl, 50 mg
thiamine-HCl, 0.5 g MgSO4, and 20 mg CaCl2 per liter of water. Cultures
were grown at 37 to an optical density (at 600 nm) of 0.6, and protein ex-
pression was induced by adding isopropyl--d-thiogalactopyranoside
(IPTG) to a concentration of 1 mM. Temperature was reduced and shaking
continued for 4 h at 30 until cells were harvested by centrifugation, and
the pellet was stored in 20 overnight. The frozen pellet (10 g) was resus-
pended in 50 ml of 50 mM potassium phosphate, pH 7.5, 10 mM imidazole,
500 mM NaCl, 5% glycerol, 1 mM benzamidine HCl, 1 mM phenylmethyl-
sulfonylfluoride (PMSF), and the suspension was stirred in the cold room
for 30 min. This suspension was then sonicated on ice with 30 pulses each
30 s using a Misonex sonicator XL. The crude lysate was then cleared by
centrifugation at 35,000 g for 30 min and the supernatant loaded onto a
column containing 3 ml of Niþ2-agarose affinity resin (Qiagen). The column
was pre-equilibrated with 30 ml of the same buffer. After all supernatant
was loaded, the column was washed overnight with 1 L of 50 mM
potassium phosphate, pH 7.5, 10 mM imidazole, 500 mM NaCl, and 5%
glycerol. Bound material was eluted from the column in 1-ml fractions with
50 mM potassium phosphate, pH 7.5, 250 mM imidazole, 100 mM NaCl,
and 5% glycerol, and fractions were analyzed for protein content and
purity by SDS-PAGE and absorption spectroscopy at 280 nm. Protein frac-
tions were pooled and concentrated using a Millipore ultrafree filtration
device and further purified by gel filtration chromatography using a Super-
dex 75 column (Pharmacia) equilibrated with 50 mM potassium phosphate,
pH 6.0, 25 mM NaCl, 1 mM sodium azide, and 0.1 mM DTT. Finally,
column fractions containing HP1 were pooled and the sample concentrated
to 1 mM for NMR studies. The HP1 chromodomain was prepared using
this protocol for all of the assays described later.
Peptide Preparation
Histone tail peptides were prepared synthetically and purified by
reverse-phase HPLC. A nonnative tyrosine residue is placed at the C-
terminus of each histone peptide to allow determination of concentration
by UV absorbance at 280 nm using the extinction coefficient e280 ¼ 1280/
M/cm. All peptides were further purified in order to remove trace organic
136 chromatin proteins [9]
domain was collected. Next, a purified and dried peptide sample was resus-
pended in the 15N-labeled chromodomain NMR sample in order to achieve
molar ratio of 1:4 protein to peptide. The [1H-15N]-HSQC spectrum of this
complex sample was collected, and compared with that of the free chromo-
domain in order to identify the sites with chemical shift changes associated
with residues on the surface of peptide binding (see Fig. 1A).
12
F. Delaglio, S. Grzesiek, G. W. Vuister, G. Zhu, J. Pfeifer, and A. Bax, J. Biomol. NMR 6,
277 (1995).
13
B. A. Johnson and R. A. Blevins, J. Biomol. NMR 4, 603 (1994).
14
S. Grzesiek and A. Bax, J. Magn. Reson. 96, 432 (1992).
15
D. Marion, P. C. Driscoll, L. E. Kay, P. T. Wingfield, A. Bax, A. M. Gronenborn, and G. M.
Clore, Biochemistry 28, 6150 (1989).
16
S. Grzesiek, S. J. Stahl, P. T. Wingfield, and A. Bax, Biochemistry 35, 10256 (1996).
[9] interaction of the chromodomain with histone peptides 137
The residues which experience the greatest change (ave > 0.15) were
then mapped onto the chromodomain surface.
17
S. Grzesiek and A. Bax, J. Am. Chem. Soc. 115, 12593 (1993).
18
C. Dhalluin, J. E. Carlson, L. Zeng, C. He, A. K. Aggarwal, and M. M. Zhou, Nature 399,
491 (1999).
19
B. P. Hudson, M. A. Martinez-Yamout, H. J. Dyson, and P. E. Wright, J. Mol. Biol. 304,
355 (2000).
20
R. H. Jacobson, A. G. Ladurner, D. S. King, and R. Tjian, Science 288, 1422 (2000).
21
D. J. Owen, P. Ornaghi, J. C. Yang, N. Lowe, P. R. Evans, P. Ballario, D. Neuhaus,
P. Filetici, and A. A. Travers, EMBO J. 19, 6141 (2000).
138 chromatin proteins [9]
Fig. 3. The methyllysine binding aromatic cage (in black) of the HP1 chromodomain
surrounding the H3 peptide methyllysine (in grey): (A) dimethyllysine and (B) trimethylly-
sine. Increasing methylation leads to a decrease in the potential to form hydrogen bonds with
a nearby ordered water molecule (in grey).
growing when 3.2 M ammonium sulfate and 0.1 M MES buffer at pH 6.1
was used as the reservoir solution. Single crystals in space group C2221
(a ¼ 33 Å, b ¼ 76 Å, c ¼ 75 Å) were obtained. The crystals were cryo-
protected in reservoir solution supplemented with 35% MPD, and flash-
frozen in liquid nitrogen for data collection. Diffraction data were collected
on a Rigaku RUH2R X-ray generator equipped with a Bruker SMART
6000 CCD detector. All data were processed and scaled with HKL2000.22
The structure was solved by molecular replacement using CNS23 and
MOLREP24 using the crystal structure of one monomer of the dimeric
Swi6 chromo shadow domain25 with all residues turned to alanines as the
search model. Chain tracing and structure refinement was carried out using
ARP/wARP,26 CNS, and O.27
Fluorescence Spectroscopy
Macromolecular associations can be monitored by either the intrinsic
tryptophan fluorescence of a protein or the fluorescence anisotropy of an
extrinsic probe attached to one of the interacting species. The use of intrin-
sic fluorescence is somewhat limited since not all macromolecular associ-
ations produce a change in intrinsic fluorescence signal. However,
intrinsic fluorescence does have an advantage in that there are no external
probes introduced to the system. Changes in fluorescence intensity upon
ligand titration can be fit to the following equation in order to attain
dissociation constants:28
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Fs ð½XT þ ½YT þ Kd Þ ð½XT þ ½YT þ KD Þ2 ð4½XT ½YT Þ
Fi ¼ ;
2½XT
(2)
where Fi is the fluorescence change produced by fitting of increasing quan-
tities of peptide, Fs the fluorescence change at saturation, [XT] the total
22
Z. Otwinowski and W. Minor, Methods Enzymol. 276, 307 (1997).
23
A. T. Brunger, P. D. Adams, G. M. Clore, W. L. DeLano, P. Gros, R. W. Grosse-Kunstleve,
J. S. Jiang, J. Kuszewski, M. Nilges, N. S. Pannu, R. J. Read, L. M. Rice, T. Simonson, and
G. L. Warren, Acta Crystallogr. D. Biol. Crystallogr. 54, 905 (1998).
24
A. Vagin and A. Teplyakov, J. Appl. Crystallogr. 80, 1022 (1997).
25
N. P. Cowieson, J. F. Partridge, R. C. Allshire, and P. J. McLaughlin, Curr. Biol. 10,
517 (2000).
26
A. Perrakis, M. Harkiolaki, K. S. Wilson, and V. S. Lamzin, Acta Crystallogr. D. Biol.
Crystallogr. 57, 1445 (2001).
27
T. A. Jones and M. Kjeldgaard, ‘‘O Version 5.9.’’ 1994.
28
K. Croce, S. J. Freedman, B. C. Furie, and B. Furie, Biochemistry 37, 16472 (1998).
140 chromatin proteins [9]
Fig. 4. Fluorescence anisotropy binding data. Binding curves for the interaction of the
HP1 chromodomain (open circles) and SuVar(3-9) chromodomain (open squares) with
fluoresceinated-H3 peptide residues 1–15 trimethylated at lysine 9. The HP1 binding assay
with the unmodified H3 peptide is plotted with filled triangles. All binding measurements
were performed at 15 in 50 mM potassium phosphate, pH 8.0, 25 mM NaCl, and 2 mM DTT.
29
S. D. Taverna, R. S. Coyne, and C. D. Allis, Cell 110, 701 (2002).
30
R. C. Cantor and P. R. Schimmel, in ‘‘Biophysical Chemistry. Part II: Techniques for the
Study of Biological Structure and Function,’’ p. 433. WH Freeman and Co., San Francisco,
1980.
31
D. M. Jameson and W. H. Sawyer, Methods Enzymol. 246, 283 (1995).
32
T. Heyduk, Y. Ma, H. Tang, and R. H. Ebright, Methods Enzymol. 274, 492 (1996).
[9] interaction of the chromodomain with histone peptides 141
Practical Considerations
When using fluorescence anisotropy to study protein-protein inter-
actions, the smaller of the two interacting proteins should carry the fluores-
cent probe. This experimental design results in the greatest difference in
fluorescence anisotropy values between the free and bound states of the
fluorophore, and thus maximal experimental sensitivity. Ideally the larger
macromolecule is serially diluted to cover a concentration range of 100-fold
below to 100-fold above the KD of the interaction being studied. Most pro-
tein-histone tail interactions observed to date occur with an affinity of
106 M. Therefore, protein samples should range from approximately
0.1 M to 1 mM for initial studies. As outlined earlier, it is very important
to keep the concentration of the fluorescently labeled peptide at least
10-fold (preferably 100-fold) less than that of the KD. Failure to do so could
result in inaccurate affinity calculations. For protein-histone tail inter-
actions, a concentration of 100 nM has been found to be a good starting
point. Concentrations of both the labeled and unlabeled proteins can then
be fine tuned after initial experiments are conducted.
Instrumentation
We have used two different instruments (Beacon 2000 and Tecan Spec-
traFluor) for this type of measurement, and obtained comparable results.
The Beacon 2000 fluorescence polarization system by PanVera has been
142 chromatin proteins [9]
Choice of Probes
Selection of fluorophores and coupling techniques are extensively de-
scribed elsewhere.33 Fluorescein is the most popular of the fluorescent
probes used in fluorescence anisotropy assays due to its high quantum yield
and long wavelengths of excitation and emission. The lifetime of the fluo-
rescein excited-state is on the order of 4 ns,30 making it possible to monitor
protein complexes of molecular weight <100 kDa. Fluorophores with
longer excited-state lifetimes should be selected in studies examining the
formation of complexes larger than 100 kDa. A wide range of conjugation
chemistries are available making it possible to conjugate fluorophores to
amines, thiols, alcohols, aldehydes, ketones, hydrazines, and carboxylic
acids.33 The fluorescent label should be placed at a position on the peptide
far enough away from the presumed interaction site to avoid interference
with binding. In our experience, placing the fluorescein label at four
residues away from binding segment of the peptide is sufficient to avoid
interference in binding affinity.
33
R. P. Haugland, ‘‘Handbook of Fluorescent Probes and Research Products.’’ Molecular
Probes Inc., Eugene, OR, 1996.
[9] interaction of the chromodomain with histone peptides 143
on the order of 108 to 103 M. This method has been exploited to examine
the interaction of the HP1 chromodomain with the H3 tail methylated at
lysine 98,11 and the TAFII250 double bromodomain with acetylated histone
H4 tail peptides.20
Small amounts of heat are either generated or absorbed by every inter-
molecular interaction. The change in enthalpy (H), binding affinity (KD),
and stoichiometry (n) of a molecular interaction can be deduced from the
measurement of such heats of interaction. These values can then be ex-
ploited to derive the changes in entropy (S), free energy (G), and heat
capacity (Cp) of binding by the following relationships:
G ¼ nRTð ln KD Þ; (6)
G ¼ H TS; (7)
@H
Cp ¼ : (8)
T
The complete thermodynamic analysis of an interaction sets ITC apart
from all other methods. A full theoretical description of ITC has been
previously reviewed and should be sought out by interested readers.34–36
ITC analysis has been used to determine the full thermodynamic
parameters of the binding of the HP1 chromodomain to H3 tail peptides
containing either trimethyllysine or dimethyllysine at position 9.11
Figure 5 shows an example of the data obtained from such an experiment
where the peptide titrated into the chromodomain sample contained tri-
methyllysine at position 9. Table I summarizes the thermodynamics of
chromodomain interaction with H3 peptides containing dimethyllysine or
trimethyllysine at position 9 as measured by ITC.
The data clearly indicate that HP1 preferentially binds to the tri-
methyllysine containing peptide over the dimethyllysine peptide. The full
TABLE I
ITC Analysis of the Binding of the HP1 Chromodomain to Methylated H3 Peptides
Me2K9 H3 6.9 0.2 9.47 0.04 1.02 6.78 0.01 2.69 0.05
Me3K9 H3 2.5 0.1 9.83 0.04 0.99 7.36 0.02 2.46 0.02
34
H. F. Fisher and N. Singh, Methods Enzymol. 259, 194 (1995).
35
L. Indyk and H. F. Fisher, Methods Enzymol. 295, 350 (1998).
36
R. L. Biltonen and N. Langerman, Methods Enzymol. 61, 287 (1979).
[9] interaction of the chromodomain with histone peptides 145
Practical Considerations
ITC measurements are extremely sensitive to differences in buffer con-
ditions; therefore, all protein and peptide solutions must be thoroughly dia-
lyzed against a suitable buffer. The two interacting species must be
dialyzed against the exact same buffer, and ideally in the same container.
This is not always possible, particularly when one species is a small peptide.
Alternatively, peptides can be exchanged into pure distilled water using a
gel filtration or desalting column before being lyophilized. The lyophilized
peptide can then be resuspended in the dialysis buffer of the protein
sample. A buffer system should be chosen that ensures both the protein
and peptide species are stable and soluble. A wide range of pH’s, buffer,
salt, and DTT concentrations are suitable for use in ITC experiments.
Phosphate buffers are particularly useful in ITC due to low ionization
enthalpy.39 All solutions must be thoroughly degassed and devoid of any
particulate matter.
The initial concentrations of the protein and peptide are critical to the
success of the experiment. The proper concentrations of macromolecules
are estimated using the following equation:
c ¼ Ka Mtot n; (9)
where c is a unit-less constant, Ka the binding association constant (Ka ¼
1/KD), Mtot the total macromolecular concentration at the beginning of
the experiment, and n is the stoichiometry of the interaction. The constant
c can be used to estimate the shape of the binding isotherm produced by
the experiment. Values of c > 5000 produce a rectangular curve with a
height equal to exactly the H of the interaction. It is not possible to accur-
ately fit the KD of an interaction when the c value is this high due to the
sharp transition of the curve at the equivalence point. In the condition of
c < 1, the curve becomes broad and featureless, making it impossible to
extract the thermodynamics of the system. Therefore, to ideally measure
the binding affinity and change in enthalpy, the c value should be above 1
and below 1000.40 However, it is not possible to accurately estimate
37
P. L. Privalov and G. I. Makhatadze, J. Mol. Biol. 213, 385 (1990).
38
G. I. Makhatadze and P. L. Privalov, J. Mol. Biol. 213, 375 (1990).
39
M. M. Pierce, C. S. Raman, and B. T. Nall, Methods 19, 213 (1999).
40
T. Wiseman, S. Williston, J. F. Brandts, and L. N. Lin, Anal. Biochem. 179, 131 (1989).
[9] interaction of the chromodomain with histone peptides 147
the c value for initial experiments unless the binding affinity and stoichiom-
etry are known fairly accurately. This may not be practical and several pre-
liminary experiments may be needed to obtain general estimates of the
stoichiometry and affinity.
Several control experiments should be conducted.39,41 First, the buffer
solution should be titrated into the cell containing the protein and the heat
observed. Additionally, the reverse experiment, where the peptide is ti-
trated into the buffer solution containing no protein should also be per-
formed if sufficient amounts of peptide are available. These control
experiments determine the heats of dilution of the system and the results
are substracted from the experimental data to obtain accurate heats.
n o12
ð1 þ ½H3 t =ðn½cd t Þ þ 1=ðnKa ½cd t ÞÞ2 4½H3 t =ðn½cd t Þ ; (10)
Acknowledgments
W.F. is a Robert Black fellow of the Damon Runyon Cancer Research Foundation. This
work was supported in part by NIH grant GM116635 to SK.
Acknowledgments
W.F. is a Robert Black fellow of the Damon Runyon Cancer Research Foundation. This
work was supported in part by NIH grant GM116635 to SK.
5
D. Nietlispach, H. R. Mott, K. M. Stott, P. R. Nielsen, A. Thiru, and E. D. Laue, in
‘‘Methods in Molecular Biology, Protein NMR Techniques. 2nd Edition: Structure
Determination of Protein Complexes by NMR’’ (K. A. Downing, ed.), Vol. 278, Humana
Press Inc., Totowa, NJ, (in press) (2004).
6
L. J. Ball, N. V. Murzina, R. W. Broadhurst, A. R. C. Raine, S. J. Archer, F. J. Stott,
A. G. Murzin, P. B. Singh, P. J. Domaille, and E. D. Laue, EMBO J. 16, 2473 (1997).
7
S. V. Brasher, B. O. Smith, R. H. Fogh, D. Nietlispach, A. Thiru, P. R. Nielsen,
R. W. Broadhurst, L. J. Ball, N. V. Murzina, and E. D. Laue, EMBO J. 19, 1587 (2000).
8
N. P. Cowieson, J. F. Partridge, R. C. Allshire, and P. J. McLaughlin, Curr. Biol. 10,
517 (2000).
9
D. A. Horita, A. V. Ivanova, A. S. Altieri, A. J. S. Klar, and R. A. Byrd, J. Mol. Biol. 307,
861 (2001).
150 chromatin proteins [10]
Fig. 1. (continued)
[10] biophysical studies of chromodomain proteins 151
10
S. Messmer, A. Franke, and R. Paro, Genes Dev. 6, 1241 (1992).
11
J. S. Platero, T. Hartnett, and J. C. Eissenberg, EMBO J. 14, 3977 (1995).
12
M. Lachner, D. O’Carrol, S. Rea, K. Mechtler, and T. Jenuwein, Nature 410, 116 (2001).
13
A. J. Bannister, P. Zegerman, J. F. Partridge, E. A. Miska, J. O. Thomas, R. C. Allshire, and
T. Kouzarides, Nature 410, 120 (2001).
14
J.-I. Nakayama, J. C. Rice, B. D. Strahl, C. D. Allis, and S. I. S. Grewal, Science 292, 110
(2001).
15
P. R. Nielsen, D. Nietlispach, H. R. Mott, J. Callaghan, A. Bannister, T. Kouzarides, A. G.
Murzin, N. V. Murzina, and E. D. Laue, Nature 416, 103 (2002).
16
S. A. Jacobs and S. Khorasanizadeh, Science 295, 2080 (2002).
17
A. Akhtar, D. Zink, and P. B. Becker, Nature 407, 405 (2000).
18
E. V. Koonin, S. Zhou, and J. C. Lucchesi, Nucleic Acids Res. 23, 4229 (1995).
19
A. Hilfiker, D. Hilfiker-Kleiner, A. Pannuti, and J. C. Lucchesi, EMBO J. 16, 2054 (1997).
[10] biophysical studies of chromodomain proteins 153
Most of the unwanted proteins will typically elute in the first 5 ml, but the
25-ml aliquot may contain some protein. Next, the protein is eluted with
25 ml of 0.5 M imidazole in buffer A. Analysis of the various fractions by
SDS gels is then needed to determine which should be pooled.
The eluted solutions contain Ni leached from the column, which may
cause precipitation during protein concentration. To avoid this we usually
pass the solutions through a Chelex resin column (Bio-Rad). The protein is
then concentrated to 5 ml in an Amicon 60-ml cell (Millipore) using a
Diaflo YM3 membrane (Millipore) and buffer exchanged on a PD-10
column (Pharmacia) into Factor Xa cutting buffer (100 mM NaCl,
50 mM Tris of pH 8.0). At this stage the OD280 is measured to estimate
the yield. (A280 ¼ 1.0 is 0.5 mg/ml for the chromodomain, 0.66 mg/ml
for the chromo shadow domain and 0.75 mg/ml for full length HP1.)
Factor Xa (Roche) is dissolved in the Factor Xa buffer and added to the
protein at between 1:50 and 1:100 (w/w). After addition of CaCl2 to
2 mM final, the sample is left overnight at 25 . (The cleaved His-tag can
then be removed by passing the sample through the Ni-column again if re-
quired.) At this stage the protein is usually sufficiently pure for most bio-
chemical applications. However, these proteins are not sufficiently stable
for structural studies and after a few weeks, storage, degradation bands
appear on SDS gels. We have found that the chromodomains can be safely
kept at 20 , whereas the shadow domain and full-length proteins tend to
aggregate after being frozen (even with flash freezing or in the presence of
glycerol). To further purify the protein, an ion exchange chromatography
step (MonoQ or MonoS (Pharmacia) columns) helps to improve the pro-
tein stability. After Factor Xa digestion we concentrate the protein solu-
tion to 5 ml using an Amicon 60-ml cell, prior to buffer exchange on the
PD-10 column into buffer B—10 mM Na phosphate (pH 7.5–8.0 for
MonoQ and pH 5.5–6.5 for MonoS). Alternatively, one can dialyze the pro-
tein in buffer B either overnight or for 2–3 h to reduce the salt concentra-
tion. A 1-ml MonoQ or MonoS column is washed/equilibrated with 5 ml
buffer B (see earlier), 10 ml buffer C (1 M NaCl in buffer B) and 5 ml
buffer B. The sample is loaded onto the column in buffer B and run on
an FPLC system at 0.5 ml/min, collecting 1-ml fractions. The column is
then washed with 10 ml of buffer A and a gradient applied from 0% to
60% C over 60 ml and from 60% to 100% C over 5 ml. The resulting
fractions are run on an SDS gel and pooled as appropriate. The shadow
domain elutes from the column in lower-salt concentration than the
chromodomain and full-length HP1 protein, which have higher surface
charge. Mass spectrometry of the samples should then be used to identify
whether any of the protein sample has become degraded or clipped at
either the N- or C-termini.
[10] biophysical studies of chromodomain proteins 155
Sample Preparation
Protein aggregation will interfere with most protein-protein interaction
assays and will decrease the quality of NMR spectra or the likelihood
of crystallization. The full-length HP1 protein and the shadow domain
usually require gel filtration to remove any aggregated protein, whereas
the chromodomains samples do not. After prolonged storage at 4 , samples
may need to be re-run on the gel filtration column to remove further
aggregates that form. For gel filtration of NMR samples, we use a
Superdex S75 column (120-ml bed volume from Pharmacia) in 100 mM
NaCl, 20 mM phosphate of pH 5.5–8.0 at 1 ml/min over 150 ml, collecting
1-ml fractions between 35 and 125 ml. The protein fractions are
pooled, concentrated, and buffer exchanged if required. To check for
any remaining aggregation, and to carry out binding assays, we run an
analogous 2.4-ml bed volume gel filtration column using a Smart system
(Pharmacia). For NMR sample preparation, the protein-containing S75
fractions were concentrated to 0.4 ml in a 10-ml Amicon cell and buffer ex-
changed into NMR buffer (10 mM Na phosphate with pH as required by
the experiment). The A280 of the solution is measured using a quartz capil-
lary (Shimadzu) with 0.5-mm diameter (sample volume of 7–10 l) to
estimate the protein concentration and 50 l D2O, 2H-DTT to 10 mM,
2
H-EDTA to 1 mM, and 0.01% NaN3 are added to the protein sample.
Finally, the sample is made up to 0.55 ml with NMR buffer.
13
Reductive Methylation and Production of C-Methyl–Labeled
Histone H3 N-Terminal Peptides
It is possible to produce peptides that are dimethylated on all the amino
groups in an unmodified substrate by reductive alkylation.20 At high pH,
and in the presence of sodium borohydride as a reducing agent, the N-ter-
minal -NH2 and the e-NH2 groups of lysine side chains will receive first
one and then another methyl group from formaldehyde. We used this
method to produce peptides that were specifically 13C-labeled in all the
methyl positions of the dimethylated amino groups. This enabled us to un-
ambiguously identify signals from the Lys-9 methyl groups in NMR spectra
of histone H3. One hundred microliters of 3.5 mM purified peptide were
mixed on ice with 900 l of freshly made 1 mg/ml NaBH4 in 200 mM NaH-
BO3 of pH 9.0 and 5 l of 20% 13C-labeled formaldehyde. The mixture
was incubated on ice for 6 min and then transferred to a new tube contain-
ing 1 mg of NaBH4 powder. A further 5 l of the formaldehyde solution
was added and the cycle was repeated eight times. The final product was
checked by mass spectrometry for uniform methylation and the absence
of cross-linking.
aromatic residues Tyr 21, Phe 45, and Trp 42. Protein fluorescence origi-
nates from the aromatic residues, Phe, Tyr, and Trp. However, the fluores-
cence of proteins containing all three aromatic amino acids is usually
dominated by tryptophan. Upon methyl lysine binding to the chromodo-
main, the microenvironment of the tryptophan changes and thus the inten-
sity of the fluorescence emission alters. This provides a mechanism,
intrinsic to the protein, by which the interaction can be studied. Steady-
state fluorescence not only identifies the occurrence of an interaction, but
can also be used to determine the affinity of the binding.
Sample Preparation
In order to perform steady-state fluorescence analysis of the chromodo-
main–histone H3 complex the protein must be present in a suitable buffer.
The buffer must maintain the protein in the correctly folded state and must
not affect the intensity of the light emitted from the tryptophan residues.
For this reason it is important to remove any imidazole left in the sample
after Ni2þ column purification. (However, it is worth noting that the pres-
ence of glutathione does not have this effect.) Thus, during the final stages
of chromodomain purification the protein was buffer exchanged into
10 mM Na phosphate of pH 7.4 and 150 mM NaCl. The solution of the
HP1 chromodomain was made up to a concentration of 3 M, as con-
firmed by amino acid analysis. The solution of the histone H3 peptide, to
which binding of the chromodomain was to be investigated, was made up
to a concentration of 1 mM within a solution of the 3 M chromodomain.
This eliminates the need to account for any dilution effect upon addition of
the peptide to the chromodomain.
Experiments
In order to perform the titrations the chromodomain solution was
placed in a 1-cm 1-cm quartz cuvette in the spectrofluorimeter (Perkin
Elmer, LS55 Luminescence Spectrometer) equipped with a thermostated
cell holder. All solutions used were previously incubated in a water bath
and all experiments were performed at 23 . After the protein solution
was transferred to the cuvette and had been inserted into the spectrofluor-
imeter, the sample was left for 10 min in order to ensure that the tempera-
ture had equilibrated. (Fluctuations in temperature are to be avoided since
they affect the thermodynamics and kinetics of the interaction, and thus the
fluorescence intensity emitted on excitation.) The sample was excited at
20
G. E. Means and R. E. Feeney, Biochemistry 7, 2192 (1968).
21
M. R. Eftink, Methods Enzymol. 278, 221 (1997).
[10] biophysical studies of chromodomain proteins 159
280 nm and emission spectra were recorded between 300 and 400 nm—ex-
citation and emission slits of 5 nm, and a scan speed of 200 s were used.
Following the initial measurement of the chromodomain alone, small ali-
quots of 0.5–2.5 l of the 1 mM histone H3 peptide solution were added se-
quentially (increasing the peptide concentration by 0.5–2.5 M at a time)
to the solution of the chromodomain within the cuvette. Upon peptide ad-
dition the solution was mixed using a plastic sterile loop and allowed to
settle prior to recording the next emission spectrum. The peak fluorescence
intensity is detected at 335 nm confirming that the chromodomain is
folded (the value for tryptophan in aqueous solution is for comparison,
348 nm). During the titration the tryptophan fluorescence intensity in-
creased on addition of the histone H3 peptide. Additions of the peptide
were then continued until no change in the fluorescence intensity was
detected and saturation of the chromodomain H3 binding site had been
achieved. Figure 2 shows a typical titration series. In this way it was
Fig. 3. Binding curve for the interaction of histone H3 tri-methylated at Lys-9 with the
chromodomain of HP1. The Kd for the interaction is 0.65 M and has been determined using
Eq. (1) (see text).
Data Analysis
The change in fluorescence intensity upon addition of an interacting
species provides a measure of the degree to which the environment of
the Trp residues of the protein under investigation become altered due to
the interaction. The resultant fluorescent intensities can be plotted as a
function of the concentration of the interacting species (Fig. 3), and the
Kd can then be determined for the interaction by least squares fitting using
the equation21:
Fobs ¼ FP þ ðFPL FP Þ=ð2½P tot Þð½P tot þ ½L tot þ Kd
sqrt ð½P tot þ ½L tot þ Kd Þ2 4½P tot ½L tot ; (1)
Results
The studies of the interaction of the chromodomain with bis-methylated
H3 peptides showed that the binding affinity of the chromodomain to the H3
peptide di-methylated at both Lys-4 and Lys-9 was very similar to that of the
analogous tri-methylated peptide. The Kd s for the chromodomain binding
to the two peptides were found to be 2.1 and 1.9 M, respectively. The Kd
for the interaction of the chromodomain with the H3 peptide tri-methylated
at only Lys-9 was found to be only slightly lower than those determined for
the interaction of the chromodomain with the bis-methylated H3 peptides,
with a Kd of 0.7 M. This implies that the methylation of Lys-4 neither con-
tributes to nor greatly interferes with the recognition of Lys-9 methylated
histone H3 by the HP1 chromodomain. The binding studies of the chromo-
domain with the H3 peptide methylated at Lys-27 showed that this peptide
binds much more weakly (Kd 12 M). Despite the sequence similarity to
the Lys-9 position, it is clear that the sequence surrounding Lys-27 does not
provide a suitable interaction partner for the chromodomain of HP1, as
was predicted from the structure.15,16
22
T. Svedberg and K. O. Pedersen, in ‘‘The Ultracentrifuge’’ (R. H. Fowler and P. Kapitza
eds.), Clarendon Press, Oxford, 1940.
162 chromatin proteins [10]
function of the radial position. For a single ideal solute the radial
dependence of the concentration can be described by22,23:
cðrÞ ¼ cðr0 Þ exp Mð1 vbar Þ!2 r2 r02 =ð2RTÞ þ K; (2)
where vbar is the partial specific volume of the solute, is the density of the
solvent (both of which can be calculated from the amino acid and buffer
composition, respectively, using Sednterp, Tom Laue and David B. Hayes,
University of New Hampshire),24 ! is the angular velocity, c(r) is the con-
centration at the radial position r in centimeters, M is the molecular weight
of the solute, R is the gas constant, and T is the temperature in degrees
Kelvin. K is a constant added to take account of imperfections in the op-
tical system. The first point included in the calculations is at r0, with c(r0)
being the concentration at this point and r is the distance from the center
of the rotor in centimetres. With a knowledge of vbar and it is possible
to derive M by least squares fitting. This expression can be expanded to
describe mixtures of molecules, the behavior of a system in equilibrium
between several species and also non-ideal behavior.22,25
Experimental Design
Normally we design our experiments such that it would be possible to
detect multimerization in a reliable manner. This essentially means per-
forming nine separate experiments to obtain enough data to allow the de-
termination of dissociation constants. This is done by preparing three
samples, containing different concentrations of the protein in the same
buffer, and performing equilibrium experiments at three different rotor
speeds. The average molecular mass of the samples should change with
starting concentration and with angular velocity for interacting systems,
but not if it is a single species.
When choosing the optimal concentration for the experiments there are
three main issues to be considered. The first is to make sure that it is pos-
sible to detect the protein in a concentration dependent manner. Second,
when looking at interactions it is necessary to work at concentrations
where all the species in the equilibrium are present in significant amounts.
Finally, it has to be taken into consideration that at high concentrations
proteins will exhibit non-ideal behavior.
The Beckman XL-I instrument that we use employs two different types
of detection: absorption (180–800 nm) and Rayleigh interference optics. It
23
J. W. Williams, K. E. Van Holde, R. L. Baldwin, and H. Fujita, Chem. Rev. 58, 715 (1958).
24
All software mentioned in this section can be obtained from the RASMB web site:
http://www.bbri.org/RASMB/
25
D. K. McRorie and P. J. Voelker, ‘‘Self-Associating Systems in the Analytical Ultracentri-
fuge.’’ Beckman Instruments Inc., Palo Alto, California.
[10] biophysical studies of chromodomain proteins 163
Sample Preparation
It is essential to use samples of high purity and quality if accurate infor-
mation about size or affinity is to be obtained. Any impurities or non-spe-
cific aggregates can contribute to the signal measured in the experiments
compromising the interpretation of the data. The best results are obtained,
if the sample is subjected to gel filtration chromatography shortly before
performing the AUC experiments to remove any non-specific higher-order
aggregates.
It is important that the buffer used as the reference in each cell corre-
sponds exactly to the solvent that holds the protein of interest. This can be
achieved in a number of ways. Typically, the fractions containing the pro-
tein after gel filtration chromatography will need to be concentrated. If this
is done by filter retention, the buffer that comes through the filter can be
used for any dilutions needed as well as for the reference buffer. Alterna-
tively, the protein sample can be dialyzed extensively against the required
final buffer and this can be used as the reference. Another method, that is
recommended by Laue,27 is to use spin columns for buffer exchange.28
The choice of buffer will also affect the result of the experiment. It is
important not to introduce substances that interfere with the detection
method used to follow the concentration of the protein. Thus, DTT cannot
be used in the buffer if protein concentrations are monitored by absorption
26
G. Ralston, ‘‘Introduction to Analytical Ultracentrifugation.’’ Beckman Instruments, Palo
Alto, California, 1993.
27
T. Laue, ‘‘Short Column Sedimentation Equilibrium Analysis for Rapid Characterization of
Macromolecules in Solution.’’ Technical Information DS-835. Beckman Coulter, Palo Alto,
California, 1992.
28
R. I. Christopherson, M. E. Jones, and L. R. Finch, Anal. Biochem. 100, 184 (1979).
164 chromatin proteins [10]
Fig. 4. AUC data for a chromodomain. Three data sets recorded using the Rayleigh
interference optics at equilibrium at 30,000 rpm are shown. The starting concentrations for the
three samples were: ()A280 ¼ 0.4, (~)A280 ¼ 0.6, and (þ)A280 ¼ 0.8. The solid lines are
calculated using the parameters obtained by fitting all nine data sets simultaneously to Eq. (2) as
described in the text. For clarity only every fifth point is plotted. The residual values left when
subtracting the calculated values of the concentration from the measured values are shown in
the top half of the graph. An even distribution around zero indicates that the data are described
well by the model. The remaining data sets were recorded at 25,000 and 34,000 rpm.
next step can be decided (Fig. 4). The shape of the residuals obtained from
the difference between the measured and calculated data can also be used as
an indication of the behavior of the system under study.25 In this case the
molecular weight calculated for the chromodomain is 10,600 Da compared
to 10,149 Da as calculated based on the amino acid sequence.
Characterizing Interactions
A ligand will alter the environment of the atoms in the amino acids that it
interacts with. This means that the chemical shift of these atoms will change
when the ligand binds and this binding can be followed by NMR spectros-
copy. From the theory of NMR it can be shown that these changes can be
categorized into one of three types—which one occurs depends on the dif-
ference in the chemical shift between the bound and the free form compared
to the exchange rate of the ligand. If the exchange is slower than the size of
the change in chemical shift between the free and bound form, then signals
corresponding to both the free and the bound forms can be observed and the
relative populations can be estimated by measuring the intensity of the two
signals. In this case the system is said to be in the slow exchange limit. This
will be the case for strong binding. In the fast exchange limit where the ex-
change is fast compared to the chemical shift difference, there will only be
one observable signal for a given nucleus with a chemical shift that is the
weighted average of the two forms. The intermediary exchange regime gives
rise to highly broadened signals that are difficult to analyze.
This means that, unless the system is in intermediate exchange, it is
possible to obtain dissociation constants for an interaction by adding the
ligand to the protein and following the change in chemical shift (or intensity)
of a well-resolved signal during the titration. One-dimensional proton spec-
tra can in principle be used. If it is possible to obtain isotope-labeled
samples, however, it is advantageous to record 2D heteronuclear experi-
ments such as a 2D 1H-15N HSQC.33–35 This spectrum will contain one signal
31
J. Cavanagh, W. J. Fairbrother, A. G. Palmer, III, and N. J. Skelton, ‘‘Protein NMR
Spectroscopy. Principles and Practice.’’ Academic Press, San Diego, 1996.
32
L. P. Mcintosh and F. W. Dahlquist, Q. Rev. Biopys. 23, 1 (1990).
[10] biophysical studies of chromodomain proteins 167
for each backbone amide proton and its attached nitrogen nucleus in the pro-
tein. The limited number of signals, combined with the improved resolution
gained from distributing the signals into two dimensions, make these spectra
some of the most convenient to analyze. The spectrometer time needed
for each experiment can be as short as 10 min, provided the sample is
sufficiently concentrated, making it feasible to perform a detailed titration.
The value of the Kd can be measured without doing any assignment of
signals in the NMR spectrum to particular atoms in the protein. If an as-
signment is available, however, it is possible to identify the amino acids
in the protein that are likely to be involved in the interaction. This is simply
done by looking to see if a signal corresponding to a given residue changes
its chemical shifts on addition of the ligand. The chemical shift is, however,
a very sensitive marker and even small structural rearrangements upon
binding can in some cases give rise to significant changes for residues far
from the actual binding site. A good example of this is the chromo shadow
domain where the addition of the PXVXL-containing peptide ligand
breaks the symmetry and alters the structure of the dimer causing two
thirds of the backbone amide signals to shift in the NMR spectra.7 In these
cases the binding site can normally still be identified by ignoring small
changes, and, if the structure is known, by only considering candidates that
are known to be surface exposed.5
33
G. Bodenhausen and D. J. Ruben, Chem. Phys. Lett. 69, 185 (1980).
34
M. Piotto, V. Saudek, and V. Sklenar, J. Biomol. NMR 2, 661 (1992).
35
V. Sklenar, M. Piotto, R. Leppik, and V. Saudek, J. Magn. Reson. Series A 102, 241 (1993).
168 chromatin proteins [10]
Fig. 5. 2D 1H-15N HSQC spectra identifying the peptide binding site. Signals from the 15N-
labeled HP1 chromodomain (10–80) without (black) and with (grey) a 2:1 excess of a
bis-(Lys-4,9) di-methylated histone H3 (1–18) peptide: (NH2 -ARTK(Me)2 QTARK
(Me)2STGGKAPGG-COOH). Spectra recorded on samples where the added peptides have
bis-tri-methylated (Lys-4,9) or tri-methylated Lys-9 give similar results.
[10] biophysical studies of chromodomain proteins 169
Fig. 6. Chemical shift differences for the 15N (open bars) and 1HN (filled bars) atoms of
the HP1 chromodomain, in the presence of histone H3, were determined from the spectra
in Fig. 5.
where pfree and pcomplex are the molar fractions of free protein and protein
in the complex, respectively, obs is the measured chemical shift, free is the
chemical shift of the free form, and complex is the chemical shift in the com-
plex. These assignments were later confirmed by sequential assignment.15
This meant that it was possible to assign a value for the change in the chem-
ical shift values () for most amino acid residues in the chromodomain
(Fig. 6). As can be seen from Fig. 7 this enabled us to confirm that the
peptide binds in the region originally proposed based on the structural
similarities between the chromodomain and Sac7d.6
A similar experiment was performed using the unmodified histone H3
peptide (data not shown). From the changes in the spectrum it can be seen
that the unmodified peptide interacts with the same binding site as the
methylated species. It is clear, however, that the interaction is very weak.
Even with a 2.5-fold excess of the peptide the average chemical shift is still
only approximately halfway between the free form and the shift for
170 chromatin proteins [10]
Fig. 7. The histone H3 binding site mapped onto the structure of the HP1
chromodomain using the chemical shift differences plotted in Fig. 6. The chromodomain
is shown as a ribbon cartoon with the peptide represented as a ball-and-stick model. The
peptide backbone is shown in white, whereas the side chain atoms are darker. Black regions of
the chromodomain ribbon show changes of more than 1 ppm. Lighter regions show smaller
changes. The figure was prepared using Molscript [P. J. Kraulis, J. Appl. Crystallogr. 24,
946 (1991)].
Introduction
Within nucleosomes, histones are targets of enzyme-catalyzed post-
translational modifications such as acetylation, phosphorylation, methyla-
tion, and ubiquitination.1–4 Dynamic modification of nucleosomes regulates
processes such as DNA synthesis, DNA repair, and transcription. Among
these modifications, reversible acetylation has emerged as a major mechan-
ism in controlling these processes. Hyperacetylation has been correlated
with transcriptionally active genes, while hypoacetylation has been corre-
lated with silenced loci. Acetylation is carried out by histone acetyltrans-
ferases, which catalyzes the transfer of the acetyl group from acetyl-CoA
to the e-amino group of lysine residues.5,6 Histone/protein deacetylases
catalyze the removal of the acetyl group from the lysine residues.7,8
Histone/protein deacetylases have been classified based on their simi-
larity to the yeast enzymes: the Rpd3-like (class I), the HDA-like (class
II), and the Silent Information Regulator 2 (Sir2)-like (class III) deacety-
lases.7–9 Both class I and class II, which are conserved in many eukaryotes
and are commonly referred to as HDACs, likely catalyze the removal of
the acetate group from the acetylated lysine residue by activation of a
water molecule for direct hydrolysis,10 generating free acetate and the de-
acetylated protein (Fig. 1). Both class I and class II have been found in
large multiprotein complexes that are recruited by DNA-binding proteins
to specific-chromatin domains. Class I and class II enzymes are sensitive to
the inhibitor Trichostatin A (TSA).7,8
1
J. Wu and M. Grunstein, Trends Biochem. Sci. 25, 619 (2000).
2
K. Luger and T. J. Richmond, Curr. Opin. Genet. Dev. 8, 140 (1998).
3
A. P. Wolffe and J. J. Hayes, Nucleic Acids Res. 27, 711 (1999).
4
P. Cheung, C. D. Allis, and P. Sassone-Corsi, Cell 103, 263 (2000).
5
M. H. Kuo and C. D. Allis, Bioessays 20, 615 (1998).
6
S. Y. Roth, J. M. Denu, and C. D. Allis, Annu. Rev. Biochem. 70, 81 (2001).
7
S. G. Gray and T. J. Ekstrom, Exp. Cell Res. 262, 75 (2001).
8
C. M. Grozinger and S. L. Schreiber, Chem. Biol. 9, 3 (2002).
9
S. Khochbin, A. Verdel, C. Lemercier, and D. Seigneurin-Berny, Curr. Opin. Genet. Dev.
11, 162 (2001).
10
M. S. Finnin, J. R. Donigian, A. Cohen, V. M. Richnon, R. A. Rifkind, P. A. Marks,
R. Breslow, and N. P. Pavletich, Nature 401, 188 (1999).
Fig. 1. Comparison of the HDAC and Sir2-like reactions. (A) Class I and class II HDACs
likely activate a water molecule for direct hydrolysis of the acetyl group from the acetylated
substrate to form deacetylated protein and free acetate as products. (B) The Sir2 family
of deacetylases catalyzes the tight coupling of NADþ cleavage and substrate deacetylation
to produce nicotinamide, deacetylated product, and O-acetyl-ADP-ribose (OAADPr).
Among the deacetylases, the Sir2 family (class III) of enzymes, also re-
ferred to as sirtuins,11,12 are unique, as they are TSA-insensitive and
NADþ-dependent.8,13–17 These enzymes are conserved in many organisms
ranging from some bacteria to humans.12 The founding member of this
[11] quantitative assays for deacetylases 173
11
R. A. Frye, Biochem. Biophys. Res. Commun. 260, 273 (1999).
12
R. A. Frye, Biochem. Biophys. Res. Commun. 273, 793 (2000).
13
S. Imai, C. M. Armstrong, M. Kaeberlein, and L. Guarente, Nature 403, 795 (2000).
14
J. S. Smith, C. B. Brachmann, I. Celic, M. A. Kenna, S. Muhammad, V. J. Starai, J. L.
Avalos, J. C. Escalante-Semerena, C. Grubmeyer, C. Wolberger, and J. D. Boeke, Proc.
Natl. Acad. Sci. USA 97, 6658 (2000).
15
J. Landry, A. Sutton, S. T. Tafrov, R. C. Heller, J. Stebbins, L. Pillus, and R. Sternglanz,
Proc. Natl. Acad. Sci. USA 97, 5807 (2000).
16
K. G. Tanner, J. Landry, R. Sternglanz, and J. M. Denu, Proc. Natl. Acad. Sci. USA 97,
14178 (2000).
17
J. M. Denu, Trends Biochem. Sci. 28, 41 (2003).
18
D. Moazed, Mol. Cell 8, 489 (2001).
19
J. C. Tanny and D. Moazed, Proc. Natl. Acad. Sci. USA 98, 415 (2001).
20
A. A. Sauve, I. Celic, J. Avalos, H. Deng, J. D. Boeke, and V. L. Schramm, Biochemistry 40,
15456 (2001).
21
M. D. Jackson and J. M. Denu, J. Biol. Chem. 277, 18535 (2002).
22
J. H. Chang, H. C. Kim, K. Y. Hwang, J. W. Lee, S. P. Jackson, S. D. Bell, and Y. Cho,
J. Biol. Chem. 277, 34489 (2002).
23
J. Landry, J. T. Slama, and R. Sternglanz, Biochem. Biophys. Res. Commun. 278, 685
(2000).
24
M. T. Borra, F. J. O’Neill, M. D. Jackson, B. Marshall, E. Verdin, K. R. Foltz, and J. M.
Denu, J. Biol. Chem. 277, 12632 (2002).
25
L. A. Rafty, M. T. Schmidt, A. L. Perraud, A. M. Scharenberg, and J. M. Denu, J. Biol.
Chem. 277, 47114 (2002).
26
J. Luo, A. Y. Nikolaev, S. Imai, D. Chen, F. Su, A. Shiloh, L. Guarente, and W. Gu, Cell
107, 137 (2001).
27
H. Vaziri, S. K. Dessain, E. Ng Eaton, S. I. Imai, R. A. Frye, T. K. Pandita, L. Guarente,
and R. A. Weinberg, Cell 107, 149 (2001).
174 chromatin proteins [11]
28
V. Muth, S. Nadaud, I. Grummt, and R. Voit, EMBO J. 20, 1353 (2001).
29
V. J. Starai, I. Celic, R. N. Cole, J. D. Boeke, and J. C. Escalante-Semerena, Science 298,
2390 (2002).
30
B. J. North, B. L. Marshall, M. T. Borra, J. M. Denu, and E. Verdin, Mol. Cell 11, 437 (2003).
31
D. Kolle, G. Brosch, T. Lechner, A. Lusser, and P. Loidl, Methods 15, 323 (1998).
32
S. Imiliani, W. Fischle, C. Van Lint, Y. Al-Abed, and E. Verdin, Proc. Natl. Acad. Sci. USA
95, 2795 (1998).
33
S. J. Darkin-Rattray, A. M. Gurnett, R. W. Myers, P. M. Dulski, T. M. Crumley, J. J.
Allocco, C. Cannova, P. T. Meinke, S. L. Colletti, M. A. Bednarek, S. B. Singh, M. A.
Goetz, A. W. Dombrowski, J. D. Polishook, and D. M. Schmatz, Proc. Natl. Acad. Sci. USA
93, 13143 (1996).
34
M. J. Hendzel, G. P. Delcuve, and J. R. Davie, J. Biol. Chem. 266, 21936 (1991).
35
R. Sendra, I. Rodrigo, M. L. Salvador, and L. Franco, Plant Mol. Biol. 11, 857 (1988).
36
A. Kervabon, J. Mery, and J. Parello, FEBS Lett. 106, 93 (1979).
37
J. Taunton, C. A. Hassig, and S. L. Schreiber, Science 272, 408 (1996).
[11] quantitative assays for deacetylases 175
38
Y. Zhang, G. LeRoy, H. P. Seelig, W. S. Lane, and D. Reinberg, Cell 95, 279 (1998).
39
Y. Zhang, N. Li, C. Caron, G. Matthias, D. Hess, S. Khochbin, and P. Matthias, EMBO
J. 22, 1168 (2003).
40
J. Landry and R. Sternglanz, Methods 31, 33 (2003).
41
B. Nare, J. J. Allocco, R. Kuningas, S. Galuska, R. W. Myers, M. A. Bednarek, and D. M.
Schmatz, Anal. Biochem. 267, 390 (1999).
176 chromatin proteins [11]
42
K. Hoffmann, G. Brosch, P. Loidl, and M. Jung, Nucleic Acids Res. 27, 2057 (1999).
43
D. Wegener, F. Wirsching, D. Riester, and A. Schwienhorst, Chem. Biol. 10, 61 (2003).
[11] quantitative assays for deacetylases 177
Deacetylase Assays
HPLC-based Deacetylase Assay. This assay relies on the separation of
substrates and products of the deacetylase reaction by a reversed-phase
HPLC. Quenched reaction mixtures are injected onto a C18 column and,
using a gradient of increased levels of organic solvent, substrates, products,
and enzyme can be resolved. This assay is applicable to all three classes
of deacetylases.
For characterization of most Sir2-like enzymes, monoacetylated H3 and
H4 peptides, corresponding to the 20 N-terminal residues of histone H3
and H4, can be used. The 110-l reactions are carried out at 37 in
þ
50 mM Tris (or phosphate), pH 7.5, with 1 mM DTT. NAD and acety-
lated peptide are mixed and pre-incubated in a 37 water bath for 5 min.
þ
Typical concentrations of NAD and acetylated peptide have ranged from
0.25 M to 1 mM; however, the range should be determined empirically, as
the Km values for these substrates may differ by orders of magnitude
depending on the Sir2 homologue (the kinetic parameter Km is discussed
in the Kinetic Analysis of the Sir2-Mediated Reaction section). The reaction
is initiated by the addition of enzyme. Although typical enzyme concentra-
tions ranging from 0.25 to 0.5 M and reaction times ranging from 1 to
15 min have been used, these parameters should be empirically deter-
mined. The concentration of enzyme and reaction time should be adjusted
such that less than 10–15% of the substrate is converted to products to
maintain steady-state conditions. The reaction is quenched by the addition
of TFA to a final concentration of 1%. Quenched samples are kept on ice
or stored in 20 if not immediately injected onto the HPLC column.
Samples are injected onto a reversed-phase HPLC column (e.g., a Vydac
C18 column, 4.6 250 mm, 201SP104) to resolve substrates and products.
A 100-l loop is typically used for the injections and the flow rate is set
to 1 ml/min. After injection, the system is run isocratically with solvent A
44
Y. Kim, K. G. Tanner, and J. M. Denu, Anal. Biochem. 280, 308 (2000).
45
G. A. Hunter and G. C. Ferreira, Anal. Biochem. 226, 221 (1995).
[11] quantitative assays for deacetylases 179
respectively. Using the HPLC-based assay, the Km and Vmax values are
2.8 0.4 M and 0.24 0.01 s1, respectively. A typical saturation curve
using the charcoal-binding assay is illustrated in Fig. 2.
TLC-based Deacetylase Assay. This assay is based on the ability to re-
solve small coenzyme metabolites using thin-layer chromatography. Anal-
yses of Sir2 activity using TLC have been previously described.15,19 Sir2
reactions are performed in the presence of radiolabeled NADþ; selection
of the label’s location depends on the desired molecule to be monitored.
To monitor the formation of nicotinamide specifically, a [14C]NADþ in
which the [14C]-label is located in the nicotinamide moiety should be used.
To monitor the formation of OAADPr or ADP-ribose, a [32P]NADþ, in
which the label is located in one of the phosphate groups, should be used.
To perform this assay, the radiolabeled NADþ and the acetylated substrate
are mixed and pre-incubated in a 37 water bath for 5 min. The reaction is
initiated by addition of the enzyme and is quenched with TFA to a final
concentration of 1%. An aliquot of 3–5 l of the reaction mixture, contain-
ing approximately 1000–10,000 CPM, is spotted onto a TLC plate (What-
man aluminum-backed silica gel plates, 250-m layer), approximately
2 cm from the bottom. Samples are spotted with at least 1 cm distance from
each other. To increase the radioactivity, it may be necessary to apply a
small volume at a time and dry the plate prior to spotting additional
volumes. The TLC plate is dried prior to placement in the chamber with
Fig. 2. NADþ saturation analysis using the charcoal-binding assay. NADþ saturation of
the Sir2-like reaction rate was performed at 37 in the presence of varying NADþ concentrations
(0.25–20 M), 25 nM HST2, a yeast Sir2 homologue, 5 M AcH3(K14), and 1 mM DTT in
50 mM Tris, pH 7.5 (at 37 ). The reaction mixture was quenched by addition of the charcoal
slurry. The product formed was quantified, and the initial velocities of the reaction at various
NADþ concentrations were plotted and fitted using the Michaelis-Menten equation.
[11] quantitative assays for deacetylases 183
Nicotinamide-Utilizing Assays
46
A. A. Sauve and V. L. Schramm, Biochemistry 42, 9249 (2003).
47
M. D. Jackson, M. T. Schmidt, N. J. Oppenheimer, and J. M. Denu, J. Biol. Chem. (in press)
(2003).
184 chromatin proteins [11]
Fig. 3. [14C]Nicotinamide-NADþ exchange reaction. The reaction was performed at 37 in
the presence of 500 M NAD þ, 300 M AcH3(K14), 0.2 M HST2, and 50 M
[14C]nicotinamide in 50 mM Tris, pH 7.5, with 1 mM DTT for the following time points:
0 min (black bar), 1 min (dark gray bar), 2 min (light gray bar), and 3 min (white bar). The
reaction was quenched with TFA to a final concentration of 1% and injected into the
reversed-phase HPLC. Radioactivity of the fractions collected from the HPLC was
determined by scintillation counting.
[11] quantitative assays for deacetylases 185
48
K. J. Bitterman, R. M. Anderson, H. Y. Cohen, M. Latorre-Esteves, and D. A. Sinclair,
J. Biol. Chem. 277, 45099 (2002).
186 chromatin proteins [11]
Sir2-like reaction, the Vmax/Km parameter includes all catalytic steps from
substrate binding up to and including the nicotinamide release step. The
kinetic parameter kcat, which reflects the maximal turnover rate of the
enzyme, can be directly obtained by dividing Vmax by the concentration
of the pure enzyme. For several Sir2-like enzymes that we have examined,
the Vmax, Km, and Vmax/Km values varied depending on the enzyme and
acetylated substrate used. We have determined Vmax values ranging from
0.2 to 0.7 s1, and depending on the enzyme-substrate pair, Km values have
ranged from <2 to 400 M. The catalytic efficiency of the enzyme also
varied depending on the acetylated substrate used, with Vmax/Km value
of 1 105 M1 s1 for good substrates and 100-fold lower for poor
substrates.
Concluding Remarks
Steady-state kinetic analyses require the accurate measurement of the
substrates consumed or products formed during the reaction. The products
of the Sir2-like reaction can be identified and quantified using the HPLC-
based, charcoal-binding, and TLC-based assays described here. These
quantitative approaches can be used to compare the reactivity of various
Sir2 homologues, to determine substrate specificity, and to explore inhibi-
tor susceptibility for Sir2-like enzymes. Though beyond the scope of this
chapter, these assays can be utilized to perform more detailed kinetic anal-
yses that may include examination of various features of the Sir2-like
chemical mechanism, identification of the critical ionizations, mutational
analysis, establishing the basic kinetic mechanism, and elucidating the
order of substrate binding and product release.49–52
Acknowledgment
We would like to acknowledge Dr. Michael D. Jackson and Manning Schmidt for helpful
discussion.
49
I. H. Segel, ‘‘Enzyme Kinetics. Behavior and Analysis of Rapid Equilibrium and Steady-
State Enzyme Systems.’’ Wiley-Interscience, New York, 1993.
50
R. A. Copeland, ‘‘Enzymes. A Practical Introduction to Structure, Mechanism, and Data
Analysis,’’ 2nd Ed., Wiley-VCH, New York, 2000.
51
A. Cornish-Bowden, ‘‘Fundamentals of Enzyme Kinetics.’’ Portland Press, London, 1995.
52
A. Fersht, ‘‘Structure and Mechanism in Protein Science. A Guide to Enzyme Catalysis and
Protein Folding.’’ Freeman, New York, 1999.
188 chromatin proteins [12]
1
R. Marmorstein, Nat. Rev. Mol. Cell. Biol. 2, 422 (2001).
2
S. L. Schreiber and B. E. Bernstein, Cell 111, 771 (2002).
3
R. L. Strausberg and S. L. Schreiber, Science 300, 294 (2003).
4
M. Yoshida, M. Kijima, M. Akita, and T. Beppu, J. Biol. Chem. 265, 17174 (1990).
5
J. Taunton, C. A. Hassig, and S. L. Schreiber, Science 272, 408 (1996).
6
J. E. Brownell, J. Zhou, T. Ranalli, R. Kobayashi, D. G. Edmondson, S. Y. Roth, and C. D.
Allis, Cell 84, 843 (1996).
7
O. D. Lau, T. K. Kundu, R. E. Soccio, S. Ait-Si-Ali, E. M. Khalil, A. Vassilev, A. P. Wolffe,
Y. Nakatani, R. G. Roeder, and P. A. Cole, Mol. Cell 5, 589 (2000).
190 chromatin proteins [12]
of the different HAT enzymes have allowed for a high degree of specificity
to be achieved in designing HAT inhibitors without intensive efforts. The
HAT inhibitory compounds reported to date fall under the category of bi-
substrate analogs. While simply designed, these compounds have proven
quite useful in the analysis of HAT function in a wide array of in vitro
and biological systems.
8
W. Zheng and P. A. Cole, Curr. Med. Chem. 9, 1187 (2002).
[12] HAT inhibitors 191
appears to reduce potency.9 Since the nature of the interaction between the
lysine-containing peptide substrate and the HAT enzymes was not known at
the outset of these efforts, a range of peptide moieties were examined. The
amino terminal tails of histones H3 and H4 were known to be likely physio-
logic acetylation targets of the HATs PCAF and GCN5 (close homologs)
and p300 and CBP (close homologs) so these sequences were selected for
the inhibitors. The length of the recognition motifs were not known so a
range of up to 20 amino acids was tested. The original synthetic approach ex-
ploited the synthesis of the peptides on solid phase with differential protec-
tion of the lysine earmarked for CoA modification.7 In these earlier efforts
the lysine was coupled to bromoacetic acid after cleavage of the peptide
from the solid support. In more recent efforts (see Fig. 3), bromoacetylation
is done while the peptide is still immobilized on the resin and the bromoace-
tamide intermediate is then cleaved from the solid phase.10,11 For Lys-CoA,
Rink amide resin is utilized, whereas for the more complex peptide-CoA
conjugates, Wang resin is used resulting in a C-terminal carboxylic acid.
Solution-phase coupling between the bromoacetamide and CoASH is gen-
erally a clean reaction and the final product purified by reversed-phase
HPLC on a C-18 column and characterized by MS. Multimilligrams of
the desired compounds have been generated in this fashion.
9
E. M. Khalil, J. De Angelis, M. Ishii, and P. A. Cole, Proc. Natl. Acad. Sci. USA 96,
12418 (1999).
10
A. N. Poux, M. Cebrat, C. M. Kim, P. A. Cole, and R. Marmorstein, Proc. Natl. Acad. Sci.
USA 99, 14065 (2002).
11
M. Cebrat, C. M. Kim, P. R. Thompson, M. Daugherty, and P. A. Cole, ‘‘Synthesis and
Analysis of Potential Prodrugs of Coenzyme A Analogs for the Inhibition of Histone
Acetyltransferase p300.’’ Bioorg. Med. Chem. 11, 3307–3313 (2003).
192 chromatin proteins [12]
Fig. 3. Synthetic schemes for (A) Lys-CoA and (B) assorted peptide-CoA conjugates
(B, Z, J represent variable amino acids that lead to the structures in Fig. 2).
have also proved to be weak inhibitors of the EsaI homolog Mof (unpub-
lished data), consistent with the nonsequential catalytic mechanisms of this
HAT family. As expected, Lys-CoA shows high potency against the p300
homolog CBP.12 In more extensive kinetic analysis, Lys-CoA was shown
to exhibit slow, tight-binding behavior with a half-life of interaction with
p300 of 4.5 min and a Ki* of 19 nM.13 Since product release appears to be
rate-limiting for p300 HAT activity, this long half-life could be rationalized
as due to a slow conformational change.
12
A. Polesskaya, I. Naguibneva, A. Duquet, S. Ait-Si-Ali, P. Robin, A. Vervish, P. Cole, and
A. Harel-Bellan, EMBO J. 20, 6816 (2001).
13
P. R. Thompson, H. Kurooka, Y. Nakatani, and P. A. Cole, J. Biol. Chem. 276, 33721 (2001).
[12] HAT inhibitors 193
TABLE I
IC50 Values for Synthetic Compounds and CoASHa
a
These measurements were made using mixed histones (33 g/ml) and acetyl-CoA
(10 M) as substrates, and full-length HAT enzymes as described in ref.7 unless
otherwise noted. NA: not assayed.
b
This measurement was made using recombinant PCAF and GCN5 catalytic domains with
50 M H3-20 peptide substrate and 10 M acetyl-CoA as described.10
c
These measurements were made using recombinant EsaI HAT domain [Y. Yan, N. A.
Barlev, R. H. Haley, S. L. Berger, and R. Marmorstein, Mol. Cell 6, 1195 (2000)] (a kind
gift of R. Marmorstein) and mixed histones (33 g/ml) and acetyl-CoA (10 M) as
substrates as described in ref.7
14
O. D. Lau, A. D. Courtney, A. Vassilev, L. A. Marzilli, R. J. Cotter, Y. Nakatani, and P. A.
Cole, J. Biol. Chem. 275, 21953 (2000).
194 chromatin proteins [12]
mechanism.14,15 Its relatively short half-life (<1 min) for binding to PCAF
was different from the behavior of Lys-CoA and p300 and was consistent
with the fact that product release does not appear to be rate-determining
for PCAF.14,15 One interesting finding was that shorter peptide-CoA conju-
gates such as H3-CoA-7 were rather weak inhibitors of PCAF and GCN5.
This result nicely correlates with studies showing that histone H3–derived
peptides considerably shorter than 20 amino acids are weak PCAF sub-
strates even though X-ray structures of the ternary complex of GCN5
reveal that most of the key interactions are mediated by the proximal
residues near Lys-14.16,17
Based on previous work on the GNAT superfamily member serotonin
N-acetyltransferase,9 a bisubstrate analog (H3-(Me)CoA-20) containing an
isopropionyl linker was synthesized10 as an epimeric mixture and shown to
be 4-fold more potent in blocking PCAF and GCN5 compared to the ori-
ginal compound (assuming only one stereoisomer to be potent). An X-ray
structure of the H3-(Me)CoA-20/GCN5 complex revealed that it was the
S-isomer (within the isopropionyl linker) of H3-(Me)CoA-20 that was
bound.10 More interestingly, the structure showed diminished interaction
between the peptide moiety of the bisubstrate analog and the enzyme
compared to the ternary complex.10 Relatively few residues of the peptide
(4–5) made contact with GCN5 and this structure seemed paradoxical in
light of the increased potency of longer peptide-CoA conjugates. However,
site-directed mutagenesis studies supported the notion that this complex
corresponds to a late catalytic intermediate in which the peptide substrate
may be dislodged in preparation for product release.10 Thus, it is likely
that the initial encounter complex between the bisubstrate analog and
GCN5, not captured by the X-ray structure, involves more extensive
peptide-enzyme interactions.
15
K. G. Tanner, M. R. Langer, and J. M. Denu, Biochemistry 39, 11961 (2000).
16
R. C. Trievel, F. Y. Li, and R. Marmorstein, Anal. Biochem. 287, 319 (2000).
17
J. R. Rojas, R. C. Trievel, J. Zhou, Y. Mo, X. Li, S. L. Berger, C. D. Allis, and
R. Marmorstein, Nature 401, 93 (1999).
[12] HAT inhibitors 195
many recruiting domains, and would not distinguish physiologic effects due
specifically to the HAT domains. Furthermore, since p300 and CBP (and
PCAF and GCN5) may show overlapping function, the more technically
demanding double knockouts may be required. Dominant negative
mutants can always be difficult to interpret because of the unwanted side
effect of protein overexpression. Moreover, none of these methods provide
the temporal control of dynamic processes or reversibility attainable with
small molecule-protein inhibitors. Thus, synthetic compounds can be quite
useful in assessing the role of HAT function. The lack of cell permeability
seen in many cases18 provides a significant limitation of the synthetic in-
hibitors described in this review. Nevertheless, both Lys-CoA and H3-
CoA-20 have been used in a series of transcriptional studies as highlighted
later. Their delivery in biological experiments can be divided into three
major approaches:
1. In vitro transcription studies with simple addition to cell-free
systems
2. Microinjection into cells
3. Concomitant addition of a lipid permeabilizing agent
18
K. Subbaramaiah, P. A. Cole, and A. J. Dannenberg, Cancer Res. 62, 2522 (2002).
19
T. K. Kundu, V. B. Palhan, Z. Y. Wang, W. An, P. A. Cole, and R. G. Roeder, Mol. Cell 6,
551 (2000).
20
W. An, V. B. Palhan, M. A. Karymov, S. H. Leuba, and R. G. Roeder, Mol. Cell 9,
811 (2002).
196 chromatin proteins [12]
(10 M), also supporting the role for p300 HAT activity in this process.21
Recently, Brady’s group and Nyborg and co-workers each demonstrated
the role of p300/CBP HAT activity as critical in transcriptional activation
by the HTLV-I Tax protein using Lys-CoA in a dosage-dependent
fashion.22,23 Nyborg’s study was especially noteworthy because it demon-
strated the role of HAT activity in transcriptional activation even within
a ‘‘tailless’’ chromatin template.23 Numerous potential p300/CBP sub-
strates were found in nuclear extracts and their acetylation was blocked
by Lys-CoA. Somewhat contrasting the findings of Roeder and co-
workers,20 Nyborg’s results hint at the potential for multiple acetylation
targets in transcriptional activation, at least in certain contexts.23
Kraus and co-workers have examined the role of p300/CBP in tran-
scriptional activation by thyroid hormone and retinoic acid receptors in
an in vitro transcriptional system.24 These studies showed a clear role for
the p300/CBP HAT activity which was inhibited by Lys-CoA. Interestingly,
the potential HAT activity of steroid receptor coactivator (SRC) proteins
was not important in transcriptional activations, although the SRCs play an
important role in recruiting p300/CBP to the hormone-regulated pro-
moters. The selective PCAF/GCN5 inhibitor H3-CoA-20 was also able,
albeit to a lesser extent, to block transcriptional activation in this system,
possibly suggesting a role for the PCAF/GCN-5 HAT activity in this
process.24
21
H. Asahara, B. Santoso, K. Du, P. A. Cole, I. Davidson, and M. Montminy, Mol. Cell Biol.
21, 7892 (2001).
22
H. Lu, C. A. Pise-Masison, T. M. Fletcher, R. L. Schiltz, A. Nagaich, M. Radonovich, G.
Hager, P. A. Cole, and J. N. Brady, Mol. Cell Biol. 22, 4450 (2002).
23
S. A. Georges, H. A. Giebler, P. A. Cole, K. Luger, P. J. Laybourn, and J. K. Nyborg, Mol.
Cell Biol. 23, 3392 (2003).
24
J. Li, K. C. Lee, P. A. Cole, J. Wong, and W. L. Kraus, Mol. Endocrinol. 17, 908 (2003).
25
M. Victor, Y. Bei, F. Gay, D., C. Mello, and Y. Shi, EMBO Rep. 3, 50 (2002).
[12] HAT inhibitors 197
contrast, H3-CoA-20 had no effect in this system. These findings were cor-
roborated by a ‘‘knock-in’’ experiment where CBP lacking a functional
HAT domain showed an identical phenotype.25
The role of p73 acetylation by p300/CBP was also probed by microin-
jection studies.26 It had been proposed that p300 (but not CBP) was re-
sponsible for acetylation of p73 in the context of DNA damage induced
by doxorubicin through a c-abl–dependent pathway. The site-specific acet-
ylation of p73 was hypothesized to be responsible for apoptosis in this set-
ting. Consistent with these proposals, microinjected Lys-CoA was able to
inhibit p73-induced apoptosis in p53 / mouse embryonic fibroblasts in
response to doxorubicin.26
Ott and co-workers27 have used Lys-CoA to investigate the role of Tat
acetylation by p300 in transcriptional activation. Microinjection of only
8 M Lys-CoA into HeLa cells caused a reduction in Tat-mediated tran-
scriptional activation of a luciferase reporter. In parallel, it was shown that
siRNA specific for p300 could abolish Tat-mediated transcriptional acti-
vation, whereas siRNA specific for CBP had no effect. Taken together,
these results argue for a specific role for p300 HAT activity in Tat-medi-
ated transcriptional activation. Connected to these findings, Ott’s group
also showed that Tat acetylation modulated its interaction with CyclinT1
and proposed that this could allow for Tat to more readily be transferred
to the elongating RNA polymerase II.
Recent studies from Wong’s group highlight the value of Lys-CoA in a
frog oocyte transcription system investigating the mechanisms of androgen
receptor and thyroid hormone receptor.28 By microinjection of Lys-CoA,
Wong demonstrated that histone acetylation was critical in transcriptional
activation and cofactor recruitment but was not key in altering DNA
supercoiling mediated by SWI/SNF.28 These studies support a sequential
mechanism for chromatin remodeling.
26
A. Costanzo, P. Merlo, N. Pediconi, M. Fulco, V. Sartorelli, P. A. Cole, G. Fontemaggi, M.
Fanciulli, L. Schiltz, G. Blandino, C. Balsano, and M. Levrero, Mol. Cell 9, 175 (2002).
27
K. Kaehlcke, A. Dorr, C. Hetzer-Egger, V. Kiermer, P. Henklein, M. Schnoelzer, E. Loret,
P. A. Cole, E. Verdin, and M. Ott, ‘‘Acetylation of Tat defines a Cyclin T1-independent
step in HIV transactivation.’’ Mol. Cell 12, 167–176 (2003).
28
Z.-Q. Huang, J. Li, L. M. Sachs, P. A. Cole, and J. Wong, EMBO J. 22, 2146 (2003).
198 chromatin proteins [12]
cell membranes, SPC was found to allow for intracellular p300/CBP HAT
activity to be blocked. Surprisingly, the enzyme remains blocked even after
immunoprecipitation and in vitro HAT assay, suggesting a very slow disso-
ciation of inhibitor. As expected, PCAF was not inhibited by Lys-CoA
under these conditions.
Using this approach, Harrel-Bellan and co-workers showed that p300/
CBP HAT activity was essential for terminal differentiation of muscle
cells.12 This contradicted a previous study arguing against a role for p300/
CBP HAT activity in muscle cell differentiation.29 The Lys-CoA inhibitor
was also valuable in providing temporal information related to muscle
differentiation. Thus, Lys-CoA blocked cell fusion and the expression of
late muscle specific markers like myosin heavy chain and myosin creatine
kinase.
Following Harel-Bellan’s approach, Medrano and co-workers success-
fully employed SPC to introduce Lys-CoA into melanocytes.30 These stud-
ies revealed that p300/CBP HAT activity was critical in preventing
senescence in melanocytes and may play a vital role in the immortalization
of these cells associated with malignant conversion. Thus, p300/CBP
inhibitors may have therapeutic value in pigmented cell neoplasia.
Summary
Selective and potent bisubstrate analog inhibitors are now available for
the PCAF/GCN5 and p300/CBP enzymes. The development of these in-
hibitors has provided for new insights into the mechanisms of histone acet-
yltransferases and the role of protein acetylation in gene regulation. They
have been particularly powerful in dissecting the role of p300/CBP HAT
activity in the context of its overall contributions to gene regulation. De-
spite their utility, several challenges still exist. No potent HAT inhibitors
have been reported for the EsaI family of HATs and no compounds can
distinguish between the close homologs p300 and CBP or PCAF and
GCN5. Moreover, increasing the cell permeability properties of existing
compounds is an important future direction. Ultimately, it will be of great
importance to know whether small molecule HAT inhibitors can make an
impact on the treatment of human disease.
29
T. A. McKinsey, C. L. Zhang, and E. N. Olson, Curr. Opin. Cell Biol. 14, 763 (2002).
30
D. Bandyopadhyay, N. A. Okan, E. Bales, L. Nascimento, P. A. Cole, and E. E. Medrano,
Cancer Res. 62, 6231 (2002).
[13] histone deacetylase inhibitors 199
Acknowledgments
We are grateful to our many collaborators whose names are mentioned in the references.
We thank Dr. R. Marmorstein for a gift of EsaI. We thank K. Miller for technical assistance.
This work was supported in part by the NIH and Ellison Medical Foundation. P.R.T. was
supported in part by a Canadian Institutes for Health Research post-doctoral fellowship.
While it has been almost four decades since the discovery that core nu-
cleosomal histones are post-translationally modified by acetylation and
methylation,1 the publications on histone modification have increased ex-
ponentially following the cloning of histone deacetylase 1 (HDAC1), and
histone acetyltransferase A (HAT A) in 1996.2,3 It is well established that
post-translational modifications, for example, acetylation, methylation, and
ubiquitination of lysine residues, phosphorylation of serine and threonine
residues, and methylation of arginine residues of the core histone tails, play
pivotal roles in regulating cellular functions involving chromatin, such as
transcription, replication, and DNA repair. Moreover, the interplay be-
tween multiple histone tail modifications has led to the histone code hy-
pothesis4,5 which states ‘‘that distinct histone modifications, on one or
more tails, act sequentially or in combination to form a ‘‘histone code,’’
that is, read by other proteins to bring about distinct downstream events.’’4
The ability to efficiently monitor changes in histone acetylation
has become increasingly important as the roles this complex event plays
in several diseases, including cancer,6 neurodegenerative diseases,7 and
autoimmune diseases,8 have become better understood. Additionally,
1
V. G. Allfrey, R. Faulkner, and A. E. Mirsky. Proc. Natl. Acad. Sci. USA 51, 786 (1964).
2
J. Taunton, C. A. Hassig, and S. L. Schreiber, Science 272, 408 (1996).
3
J. E. Brownell, J. Zhou, T. Ranalli, R. Kobayashi, D. G. Edmondson, S. Y. Roth, and C. D.
Allis, Cell 84, 843 (1996).
4
B. D. Strahl and C. D. Allis, Nature 403, 41 (2000).
5
B. M. Turner, Bioessays 22, 836 (2000).
6
P. Marks, R. A. Rifkind, V. M. Richon, R. Breslow, T. Miller, and W. K. Kelly, Nat. Rev.
Cancer 1, 194 (2001).
7
J. P. Taylor and K. H. Fischbeck, Trends Mol. Med. 8, 195 (2002).
8
N. Mishra, D. R. Brown, I. M. Olorenshaw, and G. M. Kammer, Proc. Natl. Acad. Sci. USA
98, 2628 (2001).
Acknowledgments
We are grateful to our many collaborators whose names are mentioned in the references.
We thank Dr. R. Marmorstein for a gift of EsaI. We thank K. Miller for technical assistance.
This work was supported in part by the NIH and Ellison Medical Foundation. P.R.T. was
supported in part by a Canadian Institutes for Health Research post-doctoral fellowship.
While it has been almost four decades since the discovery that core nu-
cleosomal histones are post-translationally modified by acetylation and
methylation,1 the publications on histone modification have increased ex-
ponentially following the cloning of histone deacetylase 1 (HDAC1), and
histone acetyltransferase A (HAT A) in 1996.2,3 It is well established that
post-translational modifications, for example, acetylation, methylation, and
ubiquitination of lysine residues, phosphorylation of serine and threonine
residues, and methylation of arginine residues of the core histone tails, play
pivotal roles in regulating cellular functions involving chromatin, such as
transcription, replication, and DNA repair. Moreover, the interplay be-
tween multiple histone tail modifications has led to the histone code hy-
pothesis4,5 which states ‘‘that distinct histone modifications, on one or
more tails, act sequentially or in combination to form a ‘‘histone code,’’
that is, read by other proteins to bring about distinct downstream events.’’4
The ability to efficiently monitor changes in histone acetylation
has become increasingly important as the roles this complex event plays
in several diseases, including cancer,6 neurodegenerative diseases,7 and
autoimmune diseases,8 have become better understood. Additionally,
1
V. G. Allfrey, R. Faulkner, and A. E. Mirsky. Proc. Natl. Acad. Sci. USA 51, 786 (1964).
2
J. Taunton, C. A. Hassig, and S. L. Schreiber, Science 272, 408 (1996).
3
J. E. Brownell, J. Zhou, T. Ranalli, R. Kobayashi, D. G. Edmondson, S. Y. Roth, and C. D.
Allis, Cell 84, 843 (1996).
4
B. D. Strahl and C. D. Allis, Nature 403, 41 (2000).
5
B. M. Turner, Bioessays 22, 836 (2000).
6
P. Marks, R. A. Rifkind, V. M. Richon, R. Breslow, T. Miller, and W. K. Kelly, Nat. Rev.
Cancer 1, 194 (2001).
7
J. P. Taylor and K. H. Fischbeck, Trends Mol. Med. 8, 195 (2002).
8
N. Mishra, D. R. Brown, I. M. Olorenshaw, and G. M. Kammer, Proc. Natl. Acad. Sci. USA
98, 2628 (2001).
Isolation of Histones
A common method of monitoring histone acetylation is through the
purification and evaluation of cellular histone extracts. The method men-
tioned later details the isolation of histones from cell suspensions. When
isolated tissues are used, the samples must first be homogenized to cell sus-
pensions. When using peripheral blood mononuclear cells or bone marrow
aspirates, the cells must first be isolated from the matrix by density gradient
centrifugation. Since the acetylation of histones is a reversible protein
modification, samples should be either processed immediately or flash frozen
at 80 for later processing. Although cell type–dependent, the typical
protein yield is approximately 10–100 g of histone extract per 107 cells.
9
J. S. Steffan, L. Bodai, J. Pallos, M. Poelman, A. McCampbell, B. L. Apostol, A. Kazantsev,
E. Schmidt, Y. Z. Zhu, M. Greenwald, R. Kurokawa, D. E. Housman, G. R. Jackson, J. L.
Marsh, and L. M. Thompson, Nature 413, 739 (2001).
10
E. Hockly, V. M. Richon, B. Woodman, D. L. Smith, X. Zhou, E. Rosa, K. Sathasivam,
S. Ghazi-Noori, A. Mahal, P. A. Lowden, J. S. Steffan, J. L. Marsh, L. M. Thompson, C. M.
Lewis, P. A. Marks, and G. P. Bates, Proc. Natl. Acad. Sci. USA 100, 2041 (2003).
11
N. Mishra, C. M. Reilly, D. R. Brown, P. Ruiz, and G. S. Gilkeson, J. Clin. Invest. 111, 539
(2003).
12
W. K. Kelly, O. A. O’Connor, and P. A. Marks, Expert Opin. Investig. Drugs 11, 1695 (2002).
13
R. W. Johnstone, Nat. Rev. Drug Discov. 1, 287 (2002).
[13] histone deacetylase inhibitors 201
Differential Migration
Histone acetylation results in the loss of a positive charge on the acety-
lated lysine residue. This loss of charge can be visualized as a reduction in
the migration rate of the histone subtypes (H2A, H2B, H3, and H4) during
electrophoresis through an AUT slab gel. Because each histone can be
acetylated at several potential lysine residues, AUT gels display a ladder
pattern detailing the extent of acetylation for each histone subtype. AUT
gel electrophoresis is the preferred method of analysis when evaluation
of the stoichiometry of histone acetylation is desired. The protocol
described later is modified from the procedure described by Yoshida
et al.14 Due to the position and resolution of the various histone subtypes,
this AUT gel electrophoresis protocol is especially useful for visualizing the
acetylated forms of histone H4.
14
M. Yoshida, M. Kijima, M. Akita, and T. Beppu, J. Biol. Chem. 265, 17174 (1990).
202 chromatin proteins [13]
AUT Protocol
The resolving gel [1 M acetic acid, 8 M urea, 0.5% Triton X-100,
45 mM NH4OH, 16% acrylamide] is prepared as follows:
1.7 ml glacial acetic acid
14.4 g urea
150 l 100% Triton X-100
0.9 ml saturated NH4OH
4.64 g acrylamide
0.16 g bis-acrylamide
Add H2O to 30 ml (low heat may facilitate dissolution)
200 l 10% ammonium persulfate
20 l TEMED
Layer exposed gel edge with H2O, let polymerize at 4 for 16 h
The stacking gel [1 M acetic acid, 8 M urea, 0.5% Triton X-100, 45 mM
NH4OH, 8.2% acrylamide] is prepared as follows:
1.7 ml glacial acetic acid
14.4 g urea
150 l 100% Triton X-100
0.9 ml saturated NH4OH
2.3 g acrylamide
0.16 g bis-acrylamide
Add H2O to 30 ml (low heat may facilitate dissolution)
200 l 10% APS
20 l TEMED
Let polymerize at room temperature 4 h
The gel should be pre-run at 170 V for 4 h at 4 in running buffer [0.2 M
glycine, 1 M acetic acid]. The terminals should be switched from conven-
tional electrophoresis so that the positive charged histones migrate toward
the cathode.
Normally, 1–20 g of histone extract should be used per lane depending
on the lane size and the method of protein staining to be used. The histone
samples are diluted 1:1 with AUT sample buffer [7.4 M urea, 1.4 M NH4OH,
10 mM dithiothreitol], incubated at room temperature for 5 min, and then
2.4 l of 1% pyronine Y dye in glacial acetic acid is added per 10 l of
sample. The samples are electrophoresed at 170 V for 20–30 h at 4 . The
protein bands are then visualized via Coomassie Blue or silver staining.
An example of the differential migration of histone H4 through an
AUT gel due to acetylation induced by SAHA is shown in Fig. 1.
[13] histone deacetylase inhibitors 203
Fig. 1. SAHA induces histone acetylation as visualized by Coomassie staining after AUT
gel electrophoresis. Murine erythroleukemia cells were treated with 2.5 M SAHA for 6 h.
Histone extracts were prepared and evaluated via AUT gel electrophoresis as described in
this chapter. The numbers denote the bands generated by laddering of histone H4 due to
acetylation at multiple lysine residues.
Immunological Detection
Immunological detection has become the method of choice to deter-
mine histone acetylation due to the expanding commercial availability of
antibodies that recognize key histone modifications. A plethora of anti-
bodies, both monoclonal and polyclonal, has been raised against both
native histones as well as peptides corresponding to individual acetylation
sites. The availability of these antibodies makes it possible to evaluate
global histone acetylation as well as the acetylation of specific residues on
specific histone subtypes. Detailed later are two immunological detection
methods, western blotting and immunohistochemistry, used to evaluate
histone acetylation changes.
Fig. 2. SAHA induces histone H3 acetylation in cultured T24 cells. T24 human
bladder carcinoma cells were treated with the indicated concentration of SAHA for 4 h.
The cells were then collected and histones extracted according to the described protocol.
Western blotting was performed using a polyclonal anti-acetylated histone H3 antibody from
Upstate (Ac-H3) or a polyclonal histone H3 antibody from Abcam (Total H3).
antibody overnight at 4 . The membrane is then washed and incubated with
an appropriate biotinylated secondary antibody at room temperature for
1 h. After extensive washing, the samples are developed using an avidin-
biotin-peroxidase system (Vector Laboratories, Burlingame, CA).
Examples of histone H3 acetylation detected by western blotting analysis
from histone samples isolated from SAHA treated cells are shown in Fig. 2.
Immunohistochemistry Protocol
Immunohistochemistry can be performed on cell or tissue preparations
and is useful for evaluating histone acetylation in the context of the whole
cells and tissues. For example, for samples made up of several cell types,
such as tissues, the only feasible way to discern relative histone acetylation
levels among the various cell types present is by immunohistochemistry.
For tissues and cells, samples can be fixed in buffered formalin and paraffin
embedded. For isolated cells, a less laborious method is to cytospin the
cells and fix to microscope slides.
Paraffin-embedded samples are cut into 5-M sections, deparaffinized
by conventional methods, quenched in 0.1% H2O2 to block endoge-
nous peroxidases, and subjected to antigen retrieval by boiling in 0.01 M
citric acid (pH 6.0) for 15 min. Isolated cell samples are prepared by fix-
ation with methanol/acetone prior to quenching in 0.1% H2O2. Both types
of samples are then blocked in 10% normal goat serum, followed by incu-
bation with rabbit-derived, anti-acetylated histone antibodies of interest.
After washing, the samples are incubated with a 1:1000 dilution of a bioti-
nylated goat anti-rabbit IgG secondary antibody. After extensive washing,
the samples are developed using an avidin-biotin-peroxidase system
[13] histone deacetylase inhibitors 205
Acknowledgments
We wish to thank Albert Cupo and Eddie Rosa for technical assistance.
[14] immunochemical analysis of chromatin 209
Introduction
Ever since antibodies were first used to study the organization of
histones in chromatin,1,2 immunochemical approaches have repeatedly
proved to be important tools for elucidating the structure and function of
chromatin. Accordingly, their use is constantly increasing, concomitant
with the growing repertoire of antibodies specific to nucleosomal compo-
nents, to modified histones, nucleic acids, nonhistone proteins, chromatin-
modifying enzyme complexes, and other components that modulate the
structure and activity of chromatin. Immunochemical reagents are espe-
cially versatile because antigen–antibody reactions occur under a relatively
wide range of conditions that do not markedly change the structure of
the chromatin fiber. These considerations are important in view of the dy-
namic nature of the chromatin fiber, the transient binding of nonhistone
and regulatory factors to chromatin, and the reversible posttranslational
modification of core histones.
Antibodies have a wide range of applications for the study of chromatin
structure and function, exemplified by the list in Table I. A plethora of immu-
nochemical protocols used to study histones, nucleosomes, chromatin, and
chromosomes has already been described3–7 and additional protocols are
presented in detail in the present volume of Methods in Enzymology. Ulti-
mately, all these procedures involve antigen–antibody reactions in the context
of chromatin. This chapter focuses on the major factors that govern these
reactions and that need to be considered when employing immunochemical
approaches to gain insights into questions related to chromatin.
TABLE I
Immunochemical Methods for Studies on Chromatin and
Chromatin-Interacting Components
Method Description
sites’’ placed next to the amino-terminal tail of histone H3 and the DNA,
respectively. The model illustrates that an IgG will recognize only a
small portion of the nucleosomal surface, because the antibody-binding site
comprises an area of 500–1000 Å2,8 and the cross-sectional area of the nu-
cleosome core is about 8000 Å2. As a consequence, antibody binding is
affected by changes only within the localized region in the binding site,
and nucleosomal changes that are solely distal to the binding site do not
affect antibody interactions. The model also illustrates that several anti-
bodies can simultaneously interact with a nucleosome, provided that they
are specific to different antigenic sites. Obviously, antibody accessibility to
8
S. Jones and J. M. Thornton, Proc. Natl. Acad. Sci. USA 93, 13 (1996).
[14] immunochemical analysis of chromatin 211
9
D. Goldblatt and M. Bustin, Biochemistry 14, 1689 (1975).
212 immunochemical assays of chromatin functions [14]
Immunogens
The choice of the immunogen used to elicit antibodies is the single most
important factor that determines the utility of the resulting serum. Immu-
nization with a pure, defined immunogen ensures a specific immune response
that yields antiserum with chromatin-specific antibodies. Monoclonal ap-
proaches could, in principle, circumvent the need to use defined immuno-
gens; however, the identity of an epitope recognized by the antibodies is
difficult to define. Consequently, the production of monoclonals against
complex mixtures of nuclear extracts, or even against chromatin, failed to
produce a battery of useful antibodies. We describe some immunogens used
to elicit antisera for chromatin analysis. Protocols for preparing and using
these immunogens have been described in this series and elsewhere.3,4,10
Histone Fractions
Core histones are poor immunogens and do not elicit a significant
immune response. Serum with an adequate titer of histone-specific anti-
bodies can be elicited by immunizing with purified core histone–RNA com-
plexes.10 When eliciting antibodies to purified histones, it is extremely
important to use highly purified histones that do not contain even small
amounts of highly immunogenic contaminants that could potentially elicit
a disproportionately strong immune response.
Only a small fraction of the antibodies elicited by purified histone–
RNA complexes reacts with chromatin or nucleosomes.9 As evident from
Fig. 1, access to some of the antigenic determinants in isolated histones
is sterically hindered by the DNA or by other histones. Nevertheless, be-
cause these sera have been elicited against a defined histone, they are an
excellent source of specific antibodies. Furthermore, because the mixture
of antibodies within sera is directed to a multiplicity of antigenic deter-
minants, the signal produced in Western analysis is often stronger than
that obtained with single peptide-elicited antibodies, which target a more
localized region with perhaps limited accessibility. Sera and antibodies
12
C. Maison et al., Nat. Genet. 30, 329 (2002).
214 immunochemical assays of chromatin functions [14]
Nucleosomes
Antisera to nucleosomes react well with mono- and oligonucleosomes
and linker histones, weakly with DNA and purified core histones, and not
at all with known nonhistone proteins such as HMGs.13 Digestion of nu-
cleosomes with proteases results in essentially complete loss of antigenicity
whereas DNase I digestion results in partial loss of antigenicity. The inter-
pretation of the results is obvious in light of Fig. 1. In nucleosomes, the
histones are organized into octamers whose topography is distinct from
that of isolated histones. Consequently, numerous antibodies will be
elicited against epitopes composed of surface regions near histone–histone
and histone–DNA interfaces. It is well documented that double-stranded
DNA is a poor immunogen and that few antibodies would thus be directed
against histone-free DNA regions. DNase I digestion nicks the DNA but
does not break histone–DNA contacts. Because the DNA remains at-
tached to the histone octamer, those epitopes composed of histone–DNA
contacts are intact and are still recognized by the antibodies. In contrast,
trypsin digestion cleaves the histones, thereby destroying epitopes com-
posed of histones or histone–DNA surfaces. Such proteolysis would thus
be expected to greatly reduce antigenicity.
Interestingly, antisera elicited against purified nucleosomes exhibit sig-
nificant species specificity.13 The specificity could be due to the sequence
variability of the histone amino-terminal tails, to distinct patterns of post-
translational modifications, or to both effects. By using affinity chromatog-
raphy, the sera could be a source of specific antibodies. However, the
inability to define the epitopes recognized by these antibodies seriously
limits their use for chromatin studies.
13
C. S. Tahourdin and M. Bustin, Biochemistry 19, 4387 (1980).
[14] immunochemical analysis of chromatin 215
antigenic sites recognized by the antibodies have not been clearly identi-
fied. For dehistonized chromatin, the specificity of the sera is attributed
to tissue-specific nonhistone proteins that bind tightly to DNA whereas
for chromatin preparations, the specificity could be due either to nonhis-
tone proteins or to posttranslational modifications in the histone tails.
As mentioned above, the sera can serve as analytical tools to detect chro-
matin differences between closely related cells. However, the information
gleaned from such studies is rather descriptive and thus far has not
provided new insights into chromatin structure.
Nonhistone Proteins
Antibodies are an important tool for elucidating the mechanism
whereby nonhistone proteins modify the structure and function of chroma-
tin. The choice of immunogen is guided by the same criteria as described
above; the most important consideration is rigorous purification and char-
acterization of the protein used as immunogen. Synthetic peptides corre-
sponding to an amino acid sequence deduced from genomic information
circumvent the need for protein purification and enable the production of
antibodies to modified amino acids in these proteins.
Studies with HMG proteins are a good example of the versatility of im-
munochemical techniques for analysis of the role of nonhistone proteins in
chromatin structure and function.15,16 Thus:
. Immunofluorescence, confocal immunofluorescence, immune elec-
tron microscopy, and antibody microinjections into living cells have
been used to demonstrate that HMGN proteins are associated with
transcriptionally active chromatin.17,18
. Microinjection of antibodies to HMGB1 into Pleurodeles waltlii
oocytes suggested a role for HMGB1 in transcription from
lampbrush chromosomes19 and immunofluorescence indicated that
part of the proteins are found in the cytoplasm.20
. Mobility supershift assays with specific antibodies demonstrated that
in chromatin, nucleosomes form complexes containing two molecules
of either HMGN1 or HMGN2. Furthermore, the nucleosomes
containing HMGN proteins are clustered into domains, in which
14
L. Hnilica, ‘‘Chromosomal Nonhistone Proteins.’’ CRC Press, Boca Raton, FL, 1983.
15
Y. V. Postnikov and M. Bustin, Methods Enzymol. 304, 133 (1999).
16
Y. V. Postnikov and M. Bustin, Methods Mol. Biol. 119, 303 (1999).
17
L. Einck and M. Bustin, Proc. Nat. Acad. Sci. USA 80, 6735 (1983).
18
R. Hock, F. Wilde, U. Scheer, and M. Bustin, EMBO J. 17, 6992 (1998).
19
U. Scheer, J. Sommerville, and M. Bustin, J. Cell Sci. 40, 1 (1979).
20
M. Bustin and N. Neihart, Cell 16, 181 (1979).
216 immunochemical assays of chromatin functions [14]
Nucleic Acids
A large repertoire of antibodies specific to nucleic acids and modified
bases is available and can be used for chromatin analysis. The fundamental
approaches and technical considerations for using these antibodies are
similar to those discussed above.
Antibodies
Preparation
Procedures that maximize antibody efficacy and minimize batch-
to-batch variation have been described.11 For studies with chromatin, it is
advantageous to prepare antibodies in two different species, especially if
immunolocalization studies are planned. In our laboratory we routinely
prepare antibodies in both rabbits and goats. Repeated bleedings allow ac-
cumulation of large quantities of sera (approximately 500 ml from rabbits
and more than 1 liter from goats). The spectrum of epitopes recognized
can vary from bleed to bleed. Thus, after ensuring that each bleed has an
adequate titer, we pool the various bleeds, characterize the pooled serum,
and then store it lyophilized. Under dry and cold conditions the lyophilized
[14] immunochemical analysis of chromatin 217
sera can be stored for long periods (we recover full activity after 20 years of
storage). The serum is reconstituted by adding H2O to the original volume,
and then centrifuged to remove insoluble residue.
Suka et al.7 describe the production of antibodies highly specific for dis-
tinct acetylation sites in yeast histone tails. The specificity of the sera was
determined by examining yeast strains bearing specific point mutations,
an extremely stringent test. Interestingly, the sera seem to be of sufficiently
high titer and specificity to be used directly, without additional purification.
However, in most cases, antibodies need to be purified by passage through
affinity columns containing the immunogen. The content of IgG in the
serum is on the order of 100 mg/ml and the content of specific antibodies,
which varies widely, is less than 1% of the total IgG. For anti-histone and
anti-HMG proteins, we recover about 0.2 mg of affinity-pure antibodies
per 1 ml of original serum. These solutions are stable for many years when
stored frozen.
Monoclonal antibodies can be an excellent choice, as high quantities of
these highly specific reagents can be prepared. However, caution must be
exercised because these are directed against a single epitope, which may
fortuitously resemble an epitope in a heterologous antigen, leading to
cross-reaction. Sera from patients suffering from autoimmune diseases
have served as a useful source of specific reagents.21–23 However, their rou-
tine use is limited by their limited availability, low amounts, and by the
labor–intensive process for deciphering their specificity. For most studies,
the antibody of choice is usually obtained by affinity purification from sera
elicited either by ‘‘native’’ proteins or by specific peptides.
21
R. W. Burlingame, Clin. Lab. Med. 17, 367 (1997).
22
M. Monestier, Methods 11, 36 (1997).
23
B. L. Kotzin, J. A. Lafferty, J. P. Portanova, R. L. Rubin, and E. M. Tan, J. Immunol. 133,
2554 (1984).
218 immunochemical assays of chromatin functions [14]
epitope and the strength of the contacts with the antibody. Changes in the
epitope contact residues can considerably change both the affinity and spe-
cificity of the IgG for its intended target. Indeed, it has been reported that
acetylation of Lys-9 and Lys-14 abolishes the ability of the anti-phospho
H3S10 antibody to recognize the phosphorylated epitope.24 The steric
models in Fig. 2 illustrate how significant changes may occur in the vicinity
of the phosphorylated residue on acetylating the neighboring lysines. Inser-
tion of a few atoms within this region may significantly disrupt critical
epitope–IgG contacts if they alter the positioning of phosphate to hinder
its accessibility to the IgG.
The steric considerations mentioned above for posttranslational
modifications also pertain to macromolecular interactions. This is exempli-
fied by the interaction of a nonhistone with chromatin, in the HMG Box–
DNA complex (Fig. 3A). It is obvious that potential epitopes proximate to
24
A. L. Clayton, S. Rose, M. J. Barratt, and L. C. Mahadevan, EMBO J. 19, 3714 (2000).
220 immunochemical assays of chromatin functions [14]
Concluding Remarks
The exquisite specificity of antibodies and their relative ease of
preparation render them extremely useful reagents for chromatin analysis.
They allow studies of specific components in the dynamic structure of a
living cell. To take full advantage of immunochemical approaches and
avoid pitfalls, one must carefully consider the basic features of the antibody
reaction with chromatin subunits, with their components, with nonhistones
such as HMGs, and with regulatory factors that constantly change and
remodel the structure of the chromatin fiber. The most important consider-
ations are as follows: choice of immunogen, purification of antibody, and
precautions in the experimental conditions to ensure that the immuno-
chemical reaction specifically reflects the quantity and state of the intended
antigen.
[15] antibodies against histones 221
In eukaryotic cells, all nuclear processes that require access to the DNA
template must contend with the fact that genomic DNA is packaged into
chromatin in vivo. The fundamental unit of chromatin is the nucleosome
and the definition of which is 146 bp of DNA wrapped around a histone oc-
tamer core comprising two copies each of histones H2A, H2B, H3, and
H4.1 Histone proteins are highly positively charged due to the high content
of lysine and arginine residues. In fact, many of these residues, particularly
those on the histone N-terminal tails, are subjected to post-translational
modifications such as acetylation, methylation, and ubiquitylation.2 In ad-
dition, a number of serine residues on histones are subjected to phosphor-
ylation as well. Because of the intimate contacts between histones and
DNA, these histone modifications can alter chromatin structure and in turn
play important regulatory roles in many DNA-templated processes. It was
almost 40 years ago that Allfrey et al. proposed that histone acetylation
serves to facilitate transcription3; however, it is not until recent years that
the biological roles of the different modifications on specific histone
residues are beginning to be elucidated. Just as a few examples, acetylation
of H4 at lysines 5 and 12 was found to be associated with the histone depo-
sition process, acetylation of H3 at lysines 9 and 14 as well as H4 at lysine
16 has been correlated to gene activation, and more recently, methylation
of lysine 9 of H3 has been shown to direct formation of heterochromatin
leading to gene silencing.1,4 As research into the functional roles of histone
modification continues to grow, we are beginning to understand that there
is a complex array of modified residues on the histones. Moreover, evi-
dence suggests that some histone modifications occur in combinatorial
fashions in association with specific nuclear processes.5 Given the large
number of potential combinations of modifications that can occur at the
level of the nucleosome, this has led to the proposal that histone modi-
fications collectively may form a histone or epigenetic code that specifies
1
A. P. Wolffe, in ‘‘Chromatin: Structure and Function.’’ Academic, San Diego, 1998.
2
K. E. van Holde, in ‘‘Chromatin.’’ Springer, New York, 1988.
3
V. G. Allfrey, R. Faulkner, and A. E. Mirsky, Proc. Natl. Acad. Sci. USA 51, 786 (1964).
4
Y. Zhang and D. Reinberg, Genes Dev. 15, 2343 (2001).
5
P. Cheung, C. D. Allis, and P. Sassone-Corsi, Cell 103, 263 (2000).
6
B. D. Strahl and C. D. Allis, Nature 403, 41 (2000).
7
B. M. Turner, Bioessays 22, 836 (2000).
8
T. Jenuwein and C. D. Allis, Science 293, 1074 (2001).
9
R. Lin, J. W. Leone, R. G. Cook, and C. D. Allis, J. Cell Biol. 108, 1577 (1989).
10
T. R. Hebbes, A. W. Thorne, and C. Crane-Robinson, EMBO J. 7, 1395 (1988).
11
B. M. Turner, Exp. Cell Res. 182, 206 (1989).
12
P. Jeppesen and B. M. Turner, Cell 74, 281 (1993).
13
B. A. Boggs, B. Connors, R. E. Sobel, A. C. Chinault, and C. D. Allis, Chromosoma 105,
303 (1996).
14
C. Crane-Robinson, T. R. Hebbes, A. L. Clayton, and A. W. Thorne, Methods 12, 48 (1997).
[15] antibodies against histones 223
15
P. Cheung, K. G. Tanner, W. L. Cheung, P. Sassone-Corsi, J. M. Denu, and C. D. Allis,
Mol. Cell 5, 905 (2000).
16
A. L. Clayton, S. Rose, M. J. Barratt, and L. C. Mahadevan, EMBO J. 19, 3714 (2000).
17
D. A. White, N. D. Belyaev, and B. M. Turner, Methods 19, 417 (1999).
18
L. C. Mahadevan, A. C. Willis, and M. J. Barratt, Cell 65, 775 (1991).
19
M. H. Kuo, J. E. Brownell, R. E. Sobel, T. A. Ranalli, R. G. Cook, D. G. Edmondson, S. Y.
Roth, and C. D. Allis, Nature 383, 269 (1996).
20
R. L. Schiltz, C. A. Mizzen, A. Vassilev, R. G. Cook, C. D. Allis, and Y. Nakatani, J. Biol.
Chem. 274, 1189 (1999).
224 immunochemical assays of chromatin functions [15]
ELISA
To determine whether the rabbit serum contained antibodies specific
for the di-modified H3, reactivity to the original Phos/Ac H3 peptide used
for injections, as well as to the unmodified (unmod), Ser10 phosphorylated
(Phos S10), Lys14 acetylated (Ac K14) H3 control peptides was assayed by
ELISA. A brief protocol for ELISA is as follows:
1. Dilute each peptide to the desired concentrations (ranging from 0 to
25 ng/ml) in 0.05 M sodium carbonate buffer, pH 9.6.
2. Dispense 200 l of the diluted peptide solutions into the appropriate
wells on the ELISA plate and incubate at 37 overnight.
3. Wash the wells twice with PBS-T (1 PBS, pH 7.4, 0.05% Tween
20), add 200 l PBS-T containing 1% BSA to each well, and
incubate at 37 for 1 h to block the uncoated areas of the ELISA
wells with BSA.
4. Wash the wells twice with PBS-T, add 200 l of diluted rabbit serum
(in PBS-T) to each well and incubate at 37 for 2 h. Initial tests
should be done using a range of dilution of the rabbit serum (from
1:500 to 1:2000 for this Phos/Ac H3 antiserum).
5. Wash the wells twice with PBS-T, add 200 l of diluted (1:5000)
HRP-conjugated goat anti-rabbit IgG secondary antibody to each
well, and incubate at 37 for 2 h.
6. Wash the wells twice with PBS-T, add 100 l of OPD substrate
(Sigma Fast: P-9187) per well, and let the reactions develop for
30 min.
7. At the end of the incubation period, stop reactions by adding 50 l
of 1 M H2SO4 and assay the binding of rabbit antibody to the plate-
bound test peptides by measuring the absorbance of the reactions at
492 nm using a microtiter plate reader.
Figure 1A shows a typical example of the ELISA results obtained
using a 1:1000 dilution of one of the immunized rabbit bleeds. The
results show that the rabbit serum reacted to all four peptides tested
and the same results were found with different bleeds as well as with the
[15]
antibodies against histones
Fig. 1. ELISA analyses of the crude serum from a rabbit immunized with H3 peptides phosphorylated at Ser10 and acetylated at
Lys14. (A) The crude serum was tested for reactivity to the four indicated test peptides by ELISA. Binding of antibodies to the test
peptides was measured by a colorimetric assay and quantified by absorbance at 492 nm. (B) Parallel samples of crude serum were
either mock-treated or incubated with 10 g of the indicated competing peptides prior to being tested by ELISA. The addition of
different competing peptides led to selective loss of reactivity of the antibody mixture present in the crude serum.
225
226 immunochemical assays of chromatin functions [15]
c. 10 ml of 10 mM Tris, pH 8.8
d. 10 ml of 100 mM triethanolamine, pH 11.5
e. 10 ml of 10 mM Tris, pH 7.5
3. Add 1 mg each of the Phos S10 and Ac K14 H3 peptides to the
10 ml of precipitated immunoglobulins in PBS (see step 8 of ammonium
sulfate precipitation protocol) and incubate with occasional mixing at
room temperature for 15 min.
4. Apply the immunoglobulin plus peptide mixture (step 3) over the
Phos S10/Ac K14 peptide column (step 1) and allow mixture to drip
through the column by gravity. Collect flow through and re-apply to
column one more time.
5. Wash peptide column with 3 cycles of:
a. 10 ml 1 M NaCl, 50 mM sodium acetate, pH 4.0
b. 10 ml of 1 M NaCl, 20 mM Tris, pH 8.0
6. Elute antibodies bound to peptide column with 10 ml of 100 mM
glycine, pH 2.5. Collect the eluate in four 2.5-ml fractions in separate tubes
each containing 0.5 ml of 1 M Tris, pH 8.0.
7. Wash peptide column with 10 ml of 10 mM Tris, pH 8.8.
(One can omit steps 8 and 9 if the desired antibody is known to elute in
glycine.)
8. Elute with 10 ml of 100 mM triethanolamine, pH 11.5. Collect
eluate in a tube containing 2 ml of 1 M Tris, pH 8.0.
9. Wash peptide column with 10 ml of 10 mM Tris, pH 8.8.
10. Elute one last time with 5 ml of 100 mM glycine, pH 2.5. Collect
eluate in a tube with 1 ml of 1 M Tris, pH 8.0.
11. Test the collected fractions for antibody reactivity by ELISA (see
earlier protocol).
12. Dialyze the fractions containing the desired antibody against 1
PBS until the fractions are equilibrated with 1 PBS, divide into small
aliquots and store at 80 .
21
E. Harlow and D. Lane, in ‘‘Antibodies: A Laboratory Manual.’’ CSHL Press, Cold Spring
Harbor, 1988.
[15] antibodies against histones 229
[15]
[15] antibodies against histones 231
Fig. 3. Further analysis of the specificity of the purified antibody by western blots.
Histones from mock-treated or EGF-treated cells were separated on SDS polyacrylamide
gels, transferred to PVDF membranes, and probed using a Phos H3 specific antibody or with
the affinity-purified Phos/Ac H3 antibody. Various competing peptides were added to the
antibodies prior to immunoblotting to test the antibodies’ specificities.
22
H. Wang, Z. Q. Huang, L. Xia, Q. Feng, H. Erdjument-Bromage, B. D. Strahl, S. D. Briggs,
C. D. Allis, J. Wong, P. Tempst, and Y. Zhang, Science 293, 853 (2001).
23
B. D. Strahl, S. D. Briggs, C. J. Brame, J. A. Caldwell, S. S. Koh, H. Ma, R. G. Cook,
J. Shabanowitz, D. F. Hunt, M. R. Stallcup, and C. D. Allis, Curr. Biol. 11, 996 (2001).
232 immunochemical assays of chromatin functions [15]
24
U. Preuss, G. Landsberg, and K. H. Scheidtmann, Nucleic Acids Res. 31, 878
25
M. J. Hendzel, Y. Wei, M. A. Mancini, A. Van Hooser, T. Ranalli, B. R. Brinkley, D. P.
Bazett-Jones, and C. D. Allis, Chromosoma 106, 348 (1997).
[15] antibodies against histones 233
Fig. 4. The Phos H3 and Phos/Ac H3 antibodies were tested for epitope occlusion by
ELISA tests. ELISA was performed as in Fig. 1 and 2 with the addition of a new test peptide
that corresponds to tri-modified H3 (acetylated at Lys9, phosphorylated at Ser10, and
acetylated at Lys14). The additional acetyl group on Lys9 inhibited recognition of the Phos
H3 epitope by the Phos H3 antibody but had no effect on the recognition of the Phos Ser10/
Ac Lys14 epitope by the affinity-purified antibody.
Acknowledgments
I would like to thank C. David Allis for frequent discussions and scientific support.
Portion of this work was done in the Allis lab. I would also like to thank Craig Mizzen for
excellent advice on antibody purification protocols. Finally, I would like to acknowledge
Elsevier for permission to adapt some of the figures originally published in Mol. Cell 5,
905–915 (2000) for this work.
27
M. Jaskelioff and C. L. Peterson, Nat. Cell Biol. 5, 395 (2003).
1
K. E. van Holde, ‘‘Chromatin.’’ Springer, New York, 1988.
Acknowledgments
I would like to thank C. David Allis for frequent discussions and scientific support.
Portion of this work was done in the Allis lab. I would also like to thank Craig Mizzen for
excellent advice on antibody purification protocols. Finally, I would like to acknowledge
Elsevier for permission to adapt some of the figures originally published in Mol. Cell 5,
905–915 (2000) for this work.
27
M. Jaskelioff and C. L. Peterson, Nat. Cell Biol. 5, 395 (2003).
1
K. E. van Holde, ‘‘Chromatin.’’ Springer, New York, 1988.
2
B. D. Strahl and C. D. Allis, Nature 403, 41 (2000).
3
B. M. Turner, Bioessays 22, 836 (2000).
4
T. Jenuwein and C. D. Allis, Science 293, 1074 (2001).
5
M. Lachner and T. Jenuwein, Curr. Opin. Cell Biol. 14, 286 (2002).
6
Y. Zhang and D. Reinberg, Genes Dev. 15, 2343 (2001).
7
Q. Feng, H. Wang, H. H. Ng, H. Erdjument-Bromage, P. Tempst, K. Struhl, and Y. Zhang,
Curr. Biol. 12, 1052 (2002).
8
N. Lacoste, R. T. Utley, J. M. Hunter, G. G. Poirier, and J. Cote, J. Biol. Chem. 277, 30421
(2002).
9
H. H. Ng, Q. Feng, H. Wang, H. Erdjument-Bromage, P. Tempst, Y. Zhang, and K. Struhl,
Genes Dev. 16, 1518 (2002).
10
F. van Leeuwen, P. R. Gafken, and D. E. Gottschling, Cell 109, 745 (2002).
11
J. C. Rice, K. Nishioka, K. Sarma, R. Steward, D. Reinberg, and C. D. Allis, Genes Dev. 16,
2225 (2002).
12
T. Kouzarides, Curr. Opin. Genet. Dev. 12, 198 (2002).
13
A. Vaquero, A. Loyola, and D. Reinberg, ‘‘The Constantly Changing Face of Chromatin.’’
Science’s SAGE KE: http://sageke.sciencemag.org/cgi/content/full/sageke;2003/14/re4, 2003.
14
M. A. Surani, Nature 414, 122 (2001).
15
P. A. Jones and S. B. Baylin, Nat. Rev. Genet. 3, 415 (2002).
236
immunochemical assays of chromatin functions
Fig. 1. Methyl-lysine positions in histone amino-termini and peptide design. (A) Schematic representation of
methylatable lysine positions in the histone H3 and H4 amino-termini. The H3 and H4 sequences are indicated for the
murine proteins. (B) Diagram of linear and branched peptides used to generate H3-K9 di-methylation–specific
antibodies. The amino acid (aa) numbers correspond to the positions within the H3 N-terminal tail shown in panel A.
[16]
[16] methyl-lysine histone antibodies 237
and careful controls. This is because histone lysine methylation has the
additional complexity that lysine residues can be mono-, di-, or tri-
methylated,1,16,17 with significant differences in biological output.18,19 Second,
the H3-K9 and H3-K27 positions are embedded within the same amino acid
sequence ‘‘-ARKS-’’ (see Fig. 1A), thereby allowing for considerable
cross-reactivities of the respective methyl-lysine histone antibodies.
In this chapter, we summarize our experience in the development and
characterization of rabbit polyclonal antibodies that are directed against
the methylated H3-K9 position. We compare the specificity and subnuclear
localization of several methyl H3-K9 ‘‘specific’’ antibodies that were gener-
ated in research laboratories20–22 or are available from commercial sources
(Upstate Biotech (UBI), USA and Abcam, UK). We also provide proto-
cols for peptide design, rabbit immunizations, and quality controls of
methyl-lysine histone antibodies, followed by their in vivo characterization
using indirect IF of inter- and metaphase chromatin in wild-type (wt) and
mutant mouse cells that are deficient for the Suv39h histone methyltrans-
ferases (HMTases). Our comparative analyses indicate significant discrep-
ancies in the specificity and avidity of the available methyl-lysine histone
antibodies and highlight the need for extensive quality controls, such
that experimental data can be correctly interpreted despite the exquisite
complexity of histone lysine methylation.
Results
16
W. K. Paik and S. Kim, Science 174, 114 (1971).
17
J. H. Waterborg, J. Biol. Chem. 268, 4918 (1993).
18
H. Santos-Rosa, R. Schneider, A. J. Bannister, J. Sherriff, B. E. Bernstein, N. C. Emre, S. L.
Schreiber, J. Mellor, and T. Kouzarides, Nature 419, 407 (2002).
19
A. H. Peters, S. Kubicek, L. Perez-Burgos, S. Opravil, K. Mechtler, A. Kohlmaier, A. Beyer,
M. Tachibana, Y. Shinkai, and T. Jenuwein, Mol. Cell, in press.
20
A. H. Peters, D. O’Carroll, H. Scherthan, K. Mechtler, S. Sauer, C. Schofer,
K. Weipoltshammer, M. Pagani, M. Lachner, A. Kohlmaier, S. Opravil, M. Doyle, M. Sibilia,
and T. Jenuwein, Cell 107, 323 (2001).
21
B. A. Boggs, P. Cheung, E. Heard, D. L. Spector, A. C. Chinault, and C. D. Allis, Nat.
Genet. 30, 73 (2002) (published online: 10 December 2001).
22
I. G. Cowell, R. Aucott, S. K. Mahadevaiah, P. S. Burgoyne, N. Huskisson, S. Bongiorni,
G. Prantera, L. Fanti, S. Pimpinelli, R. Wu, D. M. Gilbert, W. Shi, R. Fundele, H. Morrison,
P. Jeppesen, and P. B. Singh, Chromosoma 111, 22 (2002).
238 immunochemical assays of chromatin functions [16]
state (mono-, di-, or tri-) of the lysine of interest. To illustrate how different
peptide designs can affect antibody specificity, we have compared four
rabbit polyclonal antisera raised against various H3-K9 di-methylated
peptides (see Fig. 1B). These peptides comprise a linear octamer (aa
6–13) (UBI #07-212),21 a linear 12-mer (aa 23–34) (Abcam #ab7312), a
2-branched 11-mer (aa 5–15),19 and a 4-branched hexamer (aa 6–11).20
To assess the specificity of a given antiserum, three types of quality con-
trols are available. These comprise enzyme-linked immunosorbent assays
(ELISAS), peptide spotting analyses (dot blots), and protein blots with re-
combinant and nuclear histones. Whereas ELISAS or peptide and histone
blots are used for the initial characterization, additional protein blots with
whole nuclear extracts allow the detection of possible cross-reactivities
with other epitopes that may be present on non-histone molecules. We
have developed a comprehensive peptide blot analysis with a panel of 19
linear peptides spanning the histone H3 and H4 amino-termini and which
are specific for unmodified or mono-, di-, or tri-methylated H3-K4, H3-K9,
H3-K27, H3-K36, and H4-K20. A serial dilution of these peptides (50, 10, 2,
0.4 pmol) is spotted onto a Hybond-P membrane and binding of an anti-
body is visualized by enhanced chemiluminescence. The antibody titer
and avidity are determined by scoring binding to decreasing peptide
amounts. The four antisera described in this section were each probed at
three different dilutions, ranging from 1:200 to 1:2500.
The UBI antibody #07-212 displays a high affinity toward di-methylated
H3-K9, but also shows minor cross-reactivities (i.e., only when 50 pmol
peptides are spotted) with di-methylated H3-K4 and tri-methylated H3-
K27 (see Fig. 2). This profile is slightly different from a ‘‘golden bleed’’
batch, which lacks the minor cross-reactivity toward tri-methylated H3-
K27 (D. Allis, personal communication, data not shown). The Abcam
#ab7312 antibody also has a high preference for di-methylated H3-K9
and to a minor extent cross-reacts with di-methylated H3-K36 and
H4-K20 positions, and with tri-methylated H3-K27. This is a surprising
result, since the used antigen did not even comprise the H3-K9 residue
but was designed to generate an -dimeth H3-K27 antibody. The detected
cross-reactivities illustrate the difficulty to discriminate the highly similar
H3-K9 and H3-K27 positions. For each of these commercially available
antibodies, the most specific results were obtained at dilutions 1:1000
(UBI #07-212) and 1:500 (Abcam #ab7312).
In an attempt to increase specificity and antigenicity, we developed a
novel antiserum by using a 2-branched peptide that encompasses the di-
methylated H3-K9 position by five amino acids on either side. This -2-
dimeth H3-K9 antibody (#4679) has both a high titer (as shown by its use
at a 1:2500 dilution) and is exclusively specific for di-methylated H3-K9,
[16] methyl-lysine histone antibodies 239
Fig. 2. In vitro quality control of various H3-K9 di-methyl antibodies. Immuno-dot blots
showing reactivity of antibodies raised against linear, two- and four-branched H3-K9 di-
methylated peptide antigens (lower left corners). Antibodies were probed at various dilutions,
with the most specific reactivity and dilution shown. Dot blots contain 0, 0.4, 2, 10, and
50 pmol of linear H3 (aa 1–20; aa 19–34; aa 25–45) peptides, unmodified or mono-, di- and tri-
methylated at the K4, K9, K27, or K36 positions. In addition, a linear H4 (aa 12–31) peptide,
differentially methylated at the K20 position, was also used.
240 immunochemical assays of chromatin functions [16]
without cross-reacting with H3-K27 or any other tested position (see Fig. 2).
Using a similar strategy, we had previously designed a 4-branched peptide
spanning the di-methylated H3-K9 position by only 2–3 amino acids on
either side. The rationale for this peptide was to mimic the potential in vivo
clustering of methylated histone tails at pericentric heterochromatin, which
may be organized in a more compact nucleosome structure. The resulting
-4-dimeth H3-K9 antibody (#2233) has indeed been very useful to
detect histone methylation at pericentric heterochromatin in an Suv39h-
dependent manner.20,23 At that time, dot blot analysis with only a limited
peptide panel, including unmodified and di-methylated H3-K9 and
H3-K27 peptides, categorized this antiserum as being specific for the
di-methylated H3-K9 position (see Fig. 2).
Using the complete peptide panel in dot blot analyses, however, reveals
that the -4-dimeth H3-K9 antibody has highest affinity for tri-
methylated H3-K9 and significantly cross-reacts (i.e., with 10 pmol of
spotted peptides) with di- and tri-methylated states of all tested lysine pos-
itions (see Fig. 2), even including H1-K26 di- and tri-methylated peptides
(data not shown). The promiscuity of this antibody, which may well extend
beyond histone epitopes, can be attributed to the relatively short peptide
(only 6 aa) used to present the antigen. Despite the potential to recognize
many di- and tri-methylated lysine positions, this antiserum does not uni-
formly stain interphase chromatin but selectively decorates silenced chro-
matin regions at pericentric heterochromatin and on the inactive X
chromosome (see Figs. 3 and 4). This staining pattern has been interpreted
to be a consequence of the -4-dimeth H3-K9 antibody as having a pref-
erential affinity toward a high concentration or density of methylated lysine
residues, rather than detecting more isolated methyl epitopes.20,23 Based
on its promiscuity and potential to interact with multiple methylated lysine
positions, we would like to rename the -4-dimeth H3-K9 antibody
(# 2233) as a ‘‘multi-methyl lysine’’ antiserum.
23
C. Maison, D. Bailly, A. H. Peters, J. P. Quivy, D. Roche, A. Taddei, M. Lachner,
T. Jenuwein, and G. Almouzni, Nat. Genet. 30, 329 (2002).
[16] methyl-lysine histone antibodies 241
grade (0.16 mmol/g) are used for the entire synthesis. The peptide
synthesis is monitored by conductivity measurement.
2. The second amino acid is a DI-Fmoc-lysine-OH (Bachem,
Switzerland, cat. no. B-1610), which is used to drive the synthesis
of a 2-branched peptide via its - and e-amino groups.
[16] methyl-lysine histone antibodies 243
5. On day 35, the first bleed is withdrawn (5 ml) and analyzed by dot
blots.
6. Responding rabbits are boosted intramuscularly (rear thigh) with
200 g of KLH-peptide in PBS on day 42–45, followed by additional
boosts at subsequent timepoints.
7. On day 63, the second bleed is obtained (20 ml) and additional
bleeds (20–25 ml each) can be withdrawn every other week. The
rabbits can be maintained (more boosts are required) or exsanguin-
ated (on day 85–90; preferred procedure), yielding around 80–90 ml
of crude serum.
8. To preserve the serum, fresh 10% (w/v) sodium azide is added to a
final concentration of 0.02% (w/v). The serum is then snap-frozen in
liquid N2 and kept at 80 for long-term storage in 4-ml aliquots. A
0.5-ml working aliquot is kept at 4 .
Generation of an IgG Fraction. In most cases, it is advised to generate
an IgG fraction from the crude serum, which will improve the signal to
background ratio in IF and chromatin immunoprecipitation (ChIP) anal-
yses. Although the protein A incubation can sometimes weaken antibody
avidity, a high-titer serum will almost always gain in specificity after an
IgG fractionation.
1. To de-lipify the crude serum, mix 4 ml with an equal volume of Freon
(Aldrich, cat no. 24050-8). Vortex and centrifuge three times for
10 min at 3300 rpm at 4 . Transfer supernatant containing serum to
fresh Freon after each wash. Filter the supernatant through a 0.22-m
pore-sized filter and add 5 M NaCl to a final concentration of 1 M.
2. Apply the serum to a Protein A-Porost resin column (Applied
Biosystems) and elute antibodies in two steps, using 1.5 M MgCl2,
20 mM sodium acetate, pH 5, followed by 0.1 M glycine, pH 2.45.
3. Immediately neutralize eluted fractions with 0.1 volume of 2 M
HEPES, pH 7.9, to avoid degradation of the antibodies. Pool
positive fractions (identified by O.D. at 280 nm).
4. Dialyse the 2 ml eluate for 3 h at 4 against 2 l of PBS, followed by
an overnight incubation in fresh buffer. After dialysis, the volume
increases to 5 ml.
5. Determine antibody concentration at 200–400 nm by UV scan (for
antibody #4861, it was 1.3 mg/ml).
6. Preserve and store the dialyzed IgG fraction as described for the
crude serum.
Dot Blot Analysis. ELISA and dot blots can be used equivalently to
probe antibody specificities. We prefer dot blots, simply because the same
[16] methyl-lysine histone antibodies 245
246
Summary of Specificities for Methyl-Lysine Antibodiesa
H3-K4 Unmodified
(aa 1–20) Mono þ
Di þ þþ
Tri þ þ
H3-K9 Unmodified
(aa 1–20) Mono þþþ þþþ
Di þþþ þþþ þþþ þþ þ
Tri þþþ þþþþ þþþ þþþþ
H3-K27 Unmodified
(aa 19–34) Mono þ
Di þ
Tri þ þ þ þ þþ þþ þþþ
H3-K36 Unmodified þ þ
(aa 25–45) Mono þ
Di þ þ þþ
Tri þ þ
H4-K20 Unmodified
(aa 12–31) Mono
Di þ þþ
Tri þ þþ þþþ þ
Fraction Affinity Affinity Affinity Affinity IgG IgG IgG IgG Serum IgG
Concentration (mg/ml) 0.35 0.38 0.09 n.a. 1.7 1.7 1.3 1 n.a. 1.1
Dilution 1:1000 1:500 1:500 1:1000 1:2500 1:2500 1:2500 1:1666 1:1500 1:1500
a
Each indicated antibody has been tested at various dilutions on peptide dot blots as shown in Fig. 2. The quality control of the -2-trimeth H3-K27 antibody (#6523) is included for
comparison. The relative affinity of an antibody for a given methylated lysine residue is determined by its reactivity toward a peptide dilution series (see Fig. 2) and indicated by the number
of þ. þþþþ represents the highest antibody affinity (detecting 0.4 pmol peptide), whereas þ indicates very low-affinity (detecting 50 pmol peptide). Most of the antibodies (ab7312, UBI
07-212, #4858, #4679, #4861, #2233, Cowell et al., and #6523) have been tested by IF of mouse inter- and metaphase chromatin (see Figs. 3 and 4). Based on these IF analyses, we observed
[16]
that minor cross-reactivities (þ) can sometimes be ignored, since they do not result in significantly altered staining patterns. However, we do not know how these minor cross-reactivities
would affect readout in ChIP experiments.
n.a.: not available.
[16] methyl-lysine histone antibodies 247
25
E. Heard, C. Rougeulle, D. Arnaud, P. Avner, C. D. Allis, and D. L. Spector, Cell 107, 727
(2001).
248 immunochemical assays of chromatin functions [16]
26
M. Hampsey and D. Reinberg, Cell 113, 429 (2003).
27
K. Nishioka, J. C. Rice, K. Sarma, H. Erdjument-Bromage, J. Werner, Y. Wang, S. Chuikov,
P. Valenzuela, P. Tempst, R. Steward, J. T. Lis, C. D. Allis, and D. Reinberg, Mol. Cell 9,
1201 (2002).
28
J. Fang, Q. Feng, C. S. Ketel, H. Wang, R. Cao, L. Xia, H. Erdjument-Bromage, P. Tempst,
J. A. Simon, and Y. Zhang, Curr. Biol. 12, 1086 (2002).
29
K. Plath, J. Fang, S. K. Mlynarczyk-Evans, R. Cao, K. A. Worringer, H. Wang, C. C. de la
Cruz, A. P. Otte, B. Panning, and Y. Zhang, Science 300, 131 (2003).
30
J. Silva, W. Mak, I. Zvetkova, R. Appanah, T. B. Nesterova, Z. Webster, A. H. Peters,
T. Jenuwein, A. P. Otte, and N. Brockdorff, Dev. Cell 4, 481 (2003).
[16] methyl-lysine histone antibodies 249
Discussion
The above-mentioned examples underscore the technically complex
analysis of the functional significance for mono-, di-, and tri-methylation
of histone lysine positions in vivo. Although it is anticipated that highly
specific antibodies for each methylatable histone lysine position will be de-
veloped that can discriminate every distinct methylation state, only the mi-
nority of the currently available methyl-lysine histone antibodies fulfill all
the quality criteria and specificity controls outlined in this chapter. It is
therefore highly advised that any methyl-lysine histone antibody should
be rigorously tested in vitro and in vivo before being used in experimental
settings. Similarly, antibody companies such as Upstate Biotechnology,
USA, and Abcam, UK, would do an invaluable service to the scientific
community if they would very stringently indicate strength and weaknesses
of several of their methyl-lysine histone antibodies and, if required, stop
commercializing less-specific antibodies or sub-optimal batches.
We have taken the difficulty to discriminate between the H3-K9 and
H3-K27 positions as a quality criterion to evaluate nine different antisera
raised against the methylated H3-K9 residue (see Table I). For each anti-
serum, dot blot analyses comprising the complete peptide panel as shown
in Fig. 2 were performed. The relative affinity of an antibody for a given
methylated lysine residue was scored using reactivity toward 0.4 pmol
peptide (highest affinity; þþþþ) up to 50 pmol peptide (very low affinity;
þ). These in vitro data have then been extended and compared to in vivo
analyses by performing IF of mouse inter- and metaphase chromatin.
Although we have not done extensive IF with all antibodies, unspecific or
perturbed staining patterns became apparent if antibodies displayed cross-
reactivities in peptide dot blots ranging from high affinity (þþþþ) to
moderate/low affinity (þþ). By contrast, minor cross-reactivities (þ) can
sometimes be ignored, as they did not result in significantly altered staining
252 immunochemical assays of chromatin functions [16]
31
R. Cao, L. Wang, H. Wang, L. Xia, H. Erdjument-Bromage, P. Tempst, R. S. Jones, and
Y. Zhang, Science 298, 1039 (2002).
32
B. Czermin, R. Melfi, D. McCabe, V. Seitz, A. Imhof, and V. Pirrotta, Cell 111, 185 (2002).
33
J. Muller, C. M. Hart, N. J. Francis, M. L. Vargas, A. Sengupta, B. Wild, E. L. Miller, M. B.
O’Connor, R. E. Kingston, and J. A. Simon, Cell 111, 197 (2002).
34
A. Kuzmichev, K. Nishioka, H. Erdjument-Bromage, P. Tempst, and D. Reinberg, Genes
Dev. 16, 2893 (2002).
[16] methyl-lysine histone antibodies 253
Acknowledgments
We would like to thank Upstate Biotechnology (UBI, Lake Placid, NY) (http://
www.upstatebiotech.com/) and Abcam (Cambridge, UK) (http://www.abcam.com/) for
exchange of antibodies and critical comments to the manuscript. In particular, we would
like to acknowledge Judy Nisson, Rene Rice, Mary-Ann Jelinek, Thomas Jelinek, and Jim
Bone from Upstate Biotechnology and Darren Harper from Abcam. We also acknowledge
Gramsch Laboratories (Schwabhausen, Germany) (http://www.gramsch.de/) for excellent
antibody development. We are indebted to Judd Rice and David Allis for their contribution
and help in characterizing the new series of -2-mono-di-tri H3-K9 methyl-specific
antibodies. We thank Prim Singh for contributing the ‘‘Cowell et al.’’ antibody, and Danny
Reinberg and Bryan Turner for helpful discussions. We would like to thank Mathias
Madalinski for the protocol on peptide synthesis and Ines Steinmacher for help on mass
spectrometry analysis. Research in the laboratory of T.J. is supported by the I.M.P. through
Boehringer Ingelheim and by grants from the Vienna Economy Promotion Fund, the
European Union, and the GEN-AU initiative, which is financed by the Austrian Ministry of
Education, Science, and Culture.
35
H. H. Ng, F. Robert, R. A. Young, and K. Struhl, Mol. Cell 11, 709 (2003).
[17] analyzing antibodies 255
Histone methylation has been known to exist for over 40 years1 but the
enzymes that catalyze this reaction have remained elusive until the discovery
that Suv39H1 methylates histone H3 specifically at lysine 9.2 This discov-
ery was followed by a bevy of papers describing other methyltransferases
specific for different residues and their apparent function in vivo. Histones
are methylated at lysine as well as arginine residues. Lysines can be mono-,
di-, or tri-methylated in vivo. The sites for lysine methylation on histone
H3 are 4, 9, 27, 36, and 79 (for reviews see Zhang and Reinberg (2001),3
Turner (2002),4 Bannister et al. (2002),5 Lachner et al. (2003),6 and
Vaquero et al. (2003)7) while lysines 20 and 30 remain the sole site for
methylation on H48,9,9a and H2B,10 respectively (see Fig. 1).
Studies on the effects of these modifications on gene regulation have
been greatly facilitated by the production of antibodies ‘‘specific’’ for the
modified state. The need to carefully characterize antibodies raised against
methylated histone peptides stems from the observation by several labora-
tories that these antibodies can be promiscuous depending on several factors
such as concentration, peptide context, substrate, etc. Due to this, several
papers have been subject to scrutiny in recent months as the specificity of
the antibodies used was questionable. The results of chromatin immunopre-
cipitation (ChIP) experiments became increasingly difficult to interpret
1
K. Murray, Biochemistry 3, 10 (1964).
2
S. Rea, F. Eisenhaber, D. O’Carroll, B. D. Strahl, Z. W. Sun, M. Schmid, S. Opravil,
K. Mechtler, C. P. Ponting, C. D. Allis, and T. Jenuwein, Nature 406, 593 (2000).
3
Y. Zhang and D. Reinberg, Genes Dev. 15, 2343 (2001).
4
B. M. Turner, Cell 111, 285 (2002).
5
A. J. Bannister, R. Schneider, and T. Kouzarides, Cell 109, 801 (2002).
6
M. Lachner, R. J. O’Sullivan, and T. Jenuwein, J. Cell Sci. 116, 2117 (2003).
7
A. Vaquero, A. Loyola, and D. Reinberg, published online at http://sageke.sciencemag.org/
cgi/content/full/sageke;2003/14/re4
8
K. Nishioka, J. C. Rice, K. Sarma, H. Erdjument-Bromage, J. Werner, Y. Wang, S. Chuikov,
P. Valenzuela, P. Tempst, R. Steward, J. T. Lis, C. D. Allis, and D. Reinberg, Mol. Cell
9, 1201 (2002).
9
J. C. Rice, K. Nishioka, K. Sarma, R. Steward, D. Reinberg, and C. D. Allis, Genes Dev. 16,
2225 (2002).
9a
J. Fang, Q. Feng, C. S. Ketel, H. Wang, R. Cao, L. Xia, H. Erdjument-Bromage, P. Tempst,
J. A. Simon, and Y. Zhang, Curr. Biol. 12, 1086 (2002).
10
K. Zhang and H. Tang, J. Chromatogr. B 783, 173 (2003).
Fig. 1. A schematic representation of the various histone lysine methylation sites and their
potential degrees of methylation identified to date in vivo.
since the antibodies that had been used in these experiments were not
specific to the residue or its modified state. The ideal method to test this
would be to perform ChIP and then analyze the products that have been
immunoprecipitated by mass spectroscopy in order to confirm that the anti-
body has reacted only with the modification of interest. Since this is not
feasible, and the experiments required are too difficult to perform, initial
characterization of the antibodies and determination of optimal concentra-
tion to be used in ChIP becomes extremely important.
In this chapter, we present several parameters to be taken into consid-
eration and some useful hints for systematic characterization of antibodies
raised against methylated histone peptides. Although we have focused
on antibodies against methylated residues: H3-K4, H3-K27, and H4-K20,
the methods and procedures described herein are applicable for any anti-
body directed against the histone tail modifications, including arginine
methylation and lysine acetylation, among other modifications.
ELISA
This is the first step in antibody characterization. Analysis of the anti-
body with enzyme-linked immunosorbent assay (ELISA) is the traditional
method for characterization. The advantage of ELISA is that of all assays
described, it is the most quantitative and sensitive and the optimal concen-
tration of the antibody to be used in experiments can be ascertained, but
see later. An example of antibody characterization using ELISA is shown
in the case of polyclonal di-/tri-methyl K27 antibody (see Fig. 2). At higher
concentrations (1:300), the antibody is able to recognize both the di- and
[17] analyzing antibodies 257
Fig. 2. ELISA with polyclonal antibodies against di-/tri-methyl K27 (D.R.) shows almost
equal reactivity against both di- and tri-methyl K27 peptides at 1:300 antibody dilution. But
using 1:2700 dilution, the antibody is three times more preferential to di-methyl K27
compared to tri-methyl K27.
Fig. 3. Dot blot analyses of various antibodies raised against methylated histone peptides.
(A) Polyclonal antibodies raised against di-methyl H3 lysine K4 peptide (Upstate 07-030)
analyzed using unmodified H3 (lane 1), mono-methyl K4 (lane 2), di-methyl K4 (lane 3), and
tri-methyl K4 (lane 4). (B) Monoclonal antibodies raised against di-methyl H4K20 peptide
(D.R.) tested with unmodified H4 (lane 1), mono-methyl K20 (lane 2), di-methyl K20 (lane 3),
and tri-methyl K20 (lane 4). (C) Polyclonal antibodies raised against the branched tri-methyl
H3-K9 peptide (T.J.) analyzed for specificity using di- and tri-methyl peptides in the context of
different methylated residues. The methylation site and status of the peptides are as indicated
above the panels and the antibody used in each experiment is indicated below each panel.
histones was competed out when the western blot was performed in the
presence of the di-methyl peptide (see Fig. 4C, right panel). Immunofluo-
rescence studies also showed localization of the antibody to the nucleus
and this was lost on addition of the di-methyl K20 peptide at the time of
staining (see Fig. 4D, panel 2). Addition of non-specific peptides, in this
case di-methyl or tri-methyl H3-K27, did not affect staining (see Fig. 4D,
panels 4 and 5). Similar results were obtained using peptides containing
mono- and tri-methyl H4-K20 (data not shown).
Some problems that have been encountered in our laboratory during
the characterization of antibodies using dot blots are as follows. While
the antibody may be extremely specific for a given antigen, when the
context of the peptide is changed, the antibody is then able to recognize
even the unmodified peptide. This was seen in the case of antibodies
against methylated H3-K9, where, in the context of residues 4–15, the anti-
bodies remained specific for di- and tri-methyl K9, but the antibodies were
found to react with unmodified H3 peptide of a longer length, that is,
residues 4–32 (see Fig. 5A and B).
It is also very important to take into consideration the context of the
peptide when characterizing antibodies against the methylated H3-K9 and
H3-K27 residues. Since both of these residues are contained within a very
similar context, that is, TKQTARK9S for K9 and TKQTAARK27S for
K27, it is very likely that the antibodies will cross-react with the different
peptides. It becomes crucial then, to check every K9 antibody prepared
for cross-reactivity with K27 peptides and vice versa. Several initial reports
regarding X chromosome inactivation in mammals focused on H3-K9 meth-
ylation as the early event in X inactivation.11,12 The paper by Heard et al.11
shows that di-methylation at K9 is present on the inactive X chromosome; this
is a noteworthy point because the antibodies used in this case were the anti-
dimethyl K9 ‘‘Golden Bunny’’ antibody from David Allis’ lab and the
anti-dimethyl K9 antibodies from Upstate Biotechnology. Both of these
have been extensively characterized and have been shown to be dimethyl
K9 specific and show no cross-reactivity with other modifications. But re-
cently it has been shown without doubt that the prominent mark for the es-
tablishment of the inactive X chromosome is tri-methylation at H3-K27.13,14
11
E. Heard, C. Rougeulle, D. Arnaud, P. Avner, C. D. Allis, and D. L. Spector, Cell 107,
727 (2001).
12
A. H. Peters, J. E. Mermoud, D. O’Carroll, M. Pagani, D. Schweizer, N. Brockdorff, and
T. Jenuwein, Nat. Genet. 30, 77 (2002).
13
K. Plath, J. Fang, S. K. Mlynarczyk-Evans, R. Cao, K. A. Worringer, H. Wang, C. C. de la
Cruz, A. P. Otte, B. Panning, and Y. Zhang, Science 300, 131 (2003).
14
J. Silva, W. Mak, I. Zvetkova, R. Appanah, T. B. Nesterova, Z. Webster, A. H. Peters,
T. Jenuwein, A. P. Otte, and N. Brockdorff, Dev. Cell 4, 481 (2003).
[17] analyzing antibodies 261
Fig. 5. (A) Polyclonal antibodies raised against di-methyl H3-K9 before affinity
purification showed strong reactivity with the tri-methyl K9 peptide within the same context
(lane 4) and weak reactivity with either unmodified H3 within the same context (lane 1), di-
methyl K9 (lane 3), or the longer unmodified H3 peptide (lane 5). There was no cross-
reactivity seen with mono-methyl K9, mono-, di-, or tri-methyl K27 (lanes 2, 6, 7, and 8,
respectively). (B) After affinity purification on the di-methyl H3K9 column, the antibody is
able to recognize predominantly tri-methyl K9 (lane 5) and almost no reactivity is seen with
di-methyl K9 (lane 4). A strong signal is still seen with the longer unmodified H3 (lane 2)
while the shorter peptide showed no cross-reactivity. This antibody was used at a dilution of
1:1000. (C) ELISA analysis with the same antibody after affinity purification also confirmed
that it recognizes the tri-methyl K9 modification but not di-methyl K9, K27 peptides with
different degrees of methylation or the shorter unmodified H3 peptide. (D) Polyclonal
antibodies raised against the di-methyl H3-K27 showed specific reactivity with di- and tri-
methyl K27 (top panel), but not against methyl K9 peptides (middle panel). Competition with
the tri-methyl K27 peptide completely eliminates the signal with di-methyl K27 peptide also
(bottom panel). (E) Polyclonal antibodies raised against di-/tri-methyl H3-K27 are unable to
recognize or react to an insignificant level with di- and tri-methyl K4 and K20 peptides, as
indicated above the panel.
262 immunochemical assays of chromatin functions [17]
must be tested with recombinant and native histones run side by side on an
SDS-PAGE. A good antibody should not show any reactivity with the re-
combinant histone polypeptide (see Fig. 4C).
Sometimes, an antibody that does not exhibit reactivity with the un-
modified peptide during dot blot (Fig. 6A) or ELISA analyses will give
strong positive results with the recombinant histone polypeptide in western
blots (see Fig. 6B).
Another example is shown in Fig. 6C, in which the polyclonal antibody
generated against histone tri-methyl H4-K20 showed no reactivity with
other tri-methyl peptides during dot blot analysis. When it was then tested
by western blot analysis, it showed no reactivity with the recombinant
H4 or H3 polypeptides. But there was a strong signal seen with native H4
polypeptides and, surprisingly, with native H3 polypeptides (see Fig. 6D).
This could mean that there may be some other modification on the native
H3 that is recognized by the antibody, one that we have not tested or one
that is as yet unknown. Thus, during characterization of these antibodies,
one cannot rule out the possibility that the antibody may recognize some
modification in native histones that has not been tested or identified.
An ideal set of controls to include in the western blot characterization
is to use a series of recombinant histone polypeptides having specific
modifications that were chemically incorporated, using the technology
described by Loyola et al. in this series or that described by the Peterson
laboratory.19
Immunofluorescence
In order to test if the antibody is able to recognize its target in vivo, it
becomes essential to stain cells with this antibody. Optimal concentrations
must be determined by titration of the antibody. The knowledge that the
inactive X chromosome is enriched in H3-K27 tri-methylation facilitates
characterization of good antibodies as the inactive X signal can be identi-
fied during staining along with the loss of such a signal obtained with com-
petition using the tri-methyl peptide. For example, in Fig. 7 in which
monoclonal antibodies raised against di-/tri-methyl H3-K27 were raised,
the inactive X chromosome is detected as a strong signal in mouse embry-
onic fibroblasts (MEFs). The specificity of these antibodies in vivo was con-
firmed by competition experiments performed with methylated H3-K9
peptides. This characterization can become difficult, however, when the
exact function of the modification is unknown. One can detect heterochro-
matic or euchromatic localization depending on the modification. The most
19
M. A. Shrogen-Knaak, C. J. Fry, and C. L. Peterson, J. Biol. Chem. 278, 15744 (2003).
264 immunochemical assays of chromatin functions [17]
Fig. 6. (A) Dot blot analysis of polyclonal antibodies raised against mono-methyl H4-K20
(Abcam ab9051) shows very high specificity to the mono-methyl K20 peptide (lanes 2 and 8)
and does not recognize unmodified H4, di-methyl K20, or tri-methyl K20 peptides (lanes 1, 3,
and 4). They do not react with mono-methyl K4, K9 or K27 peptides (lanes 5, 6, and 7).
(B) Antibodies raised against mono-methyl H4-K20 (ab9051) were analyzed by western blot
analysis using recombinant nucleosomes (r) and native oligonucleosomes (N) from HeLa
cells. They recognized both unmodified and native forms of H4 equally well (lanes 1 and 2).
(C) Polyclonal antibodies raised against tri-methyl H4-K20 (D.R.) were found to be extremely
specific for the tri-methyl K20 (lane 4) and showed no reactivity with either unmodified H4, or
mono- or di-methyl K20 (lanes 1, 2, and 3, respectively) by dot blot. (D) Western blot analysis
with antibodies raised against tri-methyl H4-K20 (D.R.) shows a strong signal with the native
H4 and H3 (lane 2) but no cross-reactivity with recombinant nucleosomes (lane 1).
265
266 immunochemical assays of chromatin functions [17]
TABLE I
Upstate
Antibody Biotechnology Abcam D.R. Others
H3-K4
Mono- N/A Highly specific to N/A N/A
methyl mono-methyl
K4 only
Di- Methyl H3-K4 Methyl H3-K4 N/A N/A
methyl specific but specific but
reacts with reacts with
mono- and mono- and
di-K4 di-K4
Tri- N/A Highly specific Highly specific N/A
methyl to tri-methyl to tri-methyl
K4 only K4 only
H3-K27
Mono- N/A N/A N/A N/A
methyl
Di-methyl Highly specific N/A Methyl H3-K27 N/A
to di-methyl specific but
K27 only reacts with
di- and tri-
methyl K27
Tri- N/A N/A Methyl H3-K27 Highly specific
methyl specific but to tri-methyl
reacts with K27 only
di- and tri- (Thomas
methyl K27 Jenuwein’s
laboratory)
H4-K20
Mono- N/A Mono-methyl N/A N/A
methyl K20 specific
in dot blots
but cross-reacts
with recombinant
H4 in western
Di- Highly specific N/A Specific to N/A
methyl to di-methyl methyl K20
K20 only but cross-reacts
with mono-
methyl K20
Tri- Highly specific in Shows Shows N/A
methyl dot blot analysis cross-reactivity cross-reactivity
but shows slight with di-methyl with mono- and
reactivity to K20 di- methyl K20
trimethyl K9 in
western blot
Note: Some antibodies that are now available in various companies have been marked N/A
since they were not in circulation at the time this manuscript was prepared.
[17] analyzing antibodies 267
Concluding Remarks
Even after antibodies have been characterized by the techniques de-
scribed herein, one can perform additional experiments to confirm the anti-
body specificity. Another system that can be used for characterization of
some of these antibodies is Saccharomyces cerevisiae. This becomes espe-
cially useful when the modification to be studied exists in yeast. A yeast
strain in which the H3 and H4 chromosomal copies have been deleted
and the only source of H3 and H4 is through a plasmid-borne copy, can
be used to introduce point mutations.20 For example, H3-K4 can be mu-
tated and histones can be extracted from this strain and used for analysis
of the methyl K4 modifications. Unfortunately, not all methyl lysine modi-
fications observed in higher eukaryotes exist in S. cerevisiae such as H3-K9,
H3-K27, and H4-K20.
It has become evident that all studies that address the functional rele-
vance of the different methylated states of histones should be undertaken
only after very careful analysis of the antibody being used. The antibodies
have to be characterized by as many methods as possible and with as many
controls as possible. The number of parameters that affect the specificity of
the antibody are too many and any result that is derived from ChIP and IF
experiments must be interpreted with extreme caution.
20
W. Zhang, J. R. Bone, D. G. Edmondson, B. M. Turner, and S. Y. Roth, EMBO J. 17, 3155
(1998).
268 immunochemical assays of chromatin functions [17]
ELISA
Peptide solution was prepared at a concentration of 5 g/ml in PBS and
50 l were added to 96-well plates (Costar) and left overnight at room tem-
perature to coat the plates with antigen. All subsequent procedures were
performed at room temperature. The plates were then washed with PBST
(1 PBS with 0.05% Tween 20) twice, for 5 min each time and incubated in
blocking solution (5% bovine serum albumin in PBST) for 30 min. Plates
were then washed with PBST twice for 10 min each time and 200 l pri-
mary antibody was added at the following dilutions: 1:300, 1:900, 1:2700,
1:8100, 1:24300, 1:72900, and 1:218700. A non-specific polyclonal or mono-
clonal antibody was used as a control at 1:300 dilution. The plates were in-
cubated with primary antibody for 2 h. The plates were then washed three
times with PBST for 10 min each time. Fifty microliters of secondary anti-
body conjugated to alkaline phosphatase were added at a dilution of 1:5000
and incubated for 2 h. The plates were washed thoroughly with PBST three
times with vortexing the last time. Fifty microliters of developing solution
(Sigma Fast p-Nitrophenyl Phosphate tablet sets) were added to each well
and color was allowed to develop. The reaction was stopped by adding
50 l of 3 N NAOH. The plates were read at 405 nm on an ELISA plate
reader (Tecan).
Immunofluorescence
HeLa cells were grown overnight at 37 on cover slips. All procedures
after this were performed at room temperature. Cells were fixed with 3.7%
[18] histone methylation 269
formaldehyde in PBS for 10 min, washed twice in PBS, and then perm-
eabilized with 0.1% Triton X-100 in PBS for 10 min. Cells can then be pro-
cessed immediately for staining or can be stored in PBS for a few hours.
Cells were incubated with blocking solution (3% bovine serum albumin
in PBS) for 30 min, washed, and then primary antibody was added at
appropriate concentrations (1:30 to 1:200 dilution). The optimal concentra-
tion is determined by titration. After 1 h in solution containing primary
antibody, coverslips were washed in PBS and incubated with second-
ary antibody conjugated to rhodamine (Santa Cruz) at 1:100 dilution
for 20 min. Cover slips were washed thoroughly in PBS, three times for
10 min each time before mounting with Vectashield mounting medium
with DAPI.
For immunofluorescent staining of MEFs, refer to chapter by
J. Chaumeil et al. in this volume.
Acknowledgments
We thank Dr. Edith Heard for providing the figure with immunostaining of MEFs and
Dr. Lynne Vales for critical reading of the manuscript. We also thank Upstate Biotechnology
and Dr. Thomas Jenuwein for kindly providing us with antibodies for testing and members of
the Reinberg laboratory for helpful discussions. This work was supported by grants to D.R.
(NIH GM37120) and the Howard Hughes Medical School.
1
S. L. Berger, Curr. Opin. Genet. Dev. 12, 142 (2002).
2
B. D. Strahl and C. D. Allis, Nature (London) 403, 41 (2000).
formaldehyde in PBS for 10 min, washed twice in PBS, and then perm-
eabilized with 0.1% Triton X-100 in PBS for 10 min. Cells can then be pro-
cessed immediately for staining or can be stored in PBS for a few hours.
Cells were incubated with blocking solution (3% bovine serum albumin
in PBS) for 30 min, washed, and then primary antibody was added at
appropriate concentrations (1:30 to 1:200 dilution). The optimal concentra-
tion is determined by titration. After 1 h in solution containing primary
antibody, coverslips were washed in PBS and incubated with second-
ary antibody conjugated to rhodamine (Santa Cruz) at 1:100 dilution
for 20 min. Cover slips were washed thoroughly in PBS, three times for
10 min each time before mounting with Vectashield mounting medium
with DAPI.
For immunofluorescent staining of MEFs, refer to chapter by
J. Chaumeil et al. in this volume.
Acknowledgments
We thank Dr. Edith Heard for providing the figure with immunostaining of MEFs and
Dr. Lynne Vales for critical reading of the manuscript. We also thank Upstate Biotechnology
and Dr. Thomas Jenuwein for kindly providing us with antibodies for testing and members of
the Reinberg laboratory for helpful discussions. This work was supported by grants to D.R.
(NIH GM37120) and the Howard Hughes Medical School.
1
S. L. Berger, Curr. Opin. Genet. Dev. 12, 142 (2002).
2
B. D. Strahl and C. D. Allis, Nature (London) 403, 41 (2000).
HMTs can be divided into two groups: (i) histone lysine N-methyltrans-
ferases and (ii) histone arginine N-methyltransferases.
Histone Lysine N-Methyltransferases
All of the enzymes known to methylate lysines within histone
N-terminal tails contain a conserved methyltransferase domain termed an
SET [Su(var)3-9, Enhancer-of-zeste, Trithorax] domain.3 Comparison of
various SET domain proteins has allowed them to be classified into four
subfamilies.4 Recent evidence now strongly implicates these enzymes in
human disease, especially cancer.5,6
In vivo methylated lysines may be found in the mono-, di-, or tri-
methylated state (see Fig. 2A). It is becoming clear that these states of
methylation have differing effects with respect to chromatin structure and
transcription. For instance, di-methylation of lysine 4 of histone H3
(K4H3) is found at inactive genes as well as at active genes, whereas tri-
methylated K4H3 is found predominantly at active genes. Thus, high levels
of tri-methyl K4H3 appear to ‘‘mark’’ active genes.7
The recent solving of the crystal structure of the SET7/9 methyltrans-
ferase revealed a clue as to how the different methyl states of lysine
are regulated.8 This enzyme specifically mono-methylates K4H3. It is the
3
M. Lachner and T. Jenuwein, Curr. Opin. Cell Biol. 14, 286 (2002).
4
T. Kouzarides, Curr. Opin. Gen. Dev. 12, 198 (2002).
5
S. Varambally, S. M. Dhanasekaran, M. Zhou, T. R. Barrette, C. Kumar-Sinha, M. G.
Sanda, D. Ghosh, K. J. Pienta, R. G. Sewalt, A. P. Otte, M. A. Rubin, and A. M.
Chinnaiyan, Nature (London) 419, 572 (2002).
6
R. Schneider, A. J. Bannister, and T. Kouzarides, Trends Biochem. Sci. 27, 396 (2002).
7
H. Santos-Rosa, R. Schneider, A. J. Bannister, J. Sherriff, B. E. Bernstein, N. C. Emre, S. L.
Schreiber, J. Mellor, and T. Kouzarides, Nature (London) 419, 407 (2002).
8
B. Xiao, C. Jing, J. R. Wilson, P. A. Walker, N. Vasisht, G. Kelly, S. Howell, I. A. Taylor,
G. M. Blackburn, and S. J. Gamblin, Nature (London) 421, 652 (2003).
[18] histone methylation 271
Fig. 2. Chemical structures of methylated states of lysine (A) and arginine (B).
positioning of a tyrosine within the active site of the enzyme that prevents
the site accommodating di-methylated K4H3, which confers the specificity.
Indeed, if the tyrosine is mutated, the active site can now accommodate di-
methyl lysine and the mutated SET7/9 is now able to di- and tri-methylate
K4H3. Perhaps in vivo an allosteric change within the SET7/9 catalytic site
regulates this ‘‘switching’’ activity. Such a change could be introduced, for
example, by post-translational modification of SET7/9 or, alternatively,
protein binding partners may have an effect. Understanding the function
and interplay of these modifications requires the generation of specific
antibodies that are capable of distinguishing the various methylated forms
in vivo.7,9
9
C. Maison, D. Bailly, A. H. Peters, J. P. Quivy, D. Roche, A. Taddei, M. Lachner,
T. Jenuwein, and G. Almouzni, Nat. Genet. 30, 329 (2002).
272 immunochemical assays of chromatin functions [18]
Methods
10
Y. Zhang and D. Reinberg, Genes Dev. 15, 2343 (2001).
11
J. M. Aletta, T. R. Cimato, and M. J. Ettinger, Trends Biochem. Sci. 23, 89 (1998).
12
M. R. Stallcup, Oncogene 20, 3014 (2001).
13
D. Chen, H. Ma, H. Hong, S. S. Koh, S. M. Huang, B. T. Schurter, D. W. Aswad, and M. R.
Stallcup, Science 284, 2174 (1999).
14
G. F. Sewack, T. W. Ellis, and U Hansen, Mol. Cell. Biol. 21, 1404 (2001).
15
U. M. Bauer, S. Daujat, S. J. Nielsen, K. Nightingale, and T. Kouzarides, EMBO Rep. 3,
39 (2002).
16
S. Daujat, U.-M. Bauer, V. Shah, B. Turner, S. Berger, and T. Kouzarides, Curr. Biol. 12,
2090 (2002).
[18] histone methylation 273
Peptide Synthesis
Peptides are synthesized using Fmoc chemistry17 on an Applied Biosys-
tems 9050þ or Pioneer automatic peptide synthesizer with customized
17
E. Atherton and R. C. Sheppard, eds., ‘‘Solid Phase Peptide Synthesis.’’ IRL Press (Oxford
University) 1989.
274 immunochemical assays of chromatin functions [18]
18
G. E. Means and R. E. Feeney, Biochemistry 7, 2192 (1968).
276 immunochemical assays of chromatin functions [18]
Inoculating Rabbits
Peptides containing methylated residues are usually coupled to
KLH carrier protein using standard techniques. However, if the immuno-
gen is a chemically methylated protein, then it is used directly. A typical
immunization schedule is outlined later. Before initiating the immuniza-
tion, pre-immune bleeds are taken from the rabbit to identify existing
immune responses. The first immunization (week 0) injection contains
200 g of immunogen per rabbit emulsified in complete Freund’s adjuvant
and this is then followed by two further immunizations (weeks 3 and 6) of
100 g of immunogen per rabbit in incomplete Freund’s adjuvant. Our in-
jection procedure typically involves four subcutaneous site injections per
rabbit per immunization. Two bleeds at weeks 7 and 9 are then taken,
followed by a further immunization in incomplete Freund’s adjuvant.
Two further bleeds are taken at weeks 11 and 13 and finally exsanguination
of the rabbit occurs at week 14. The blood is centrifuged to isolate the sera.
The level and specificity of immune response is then determined using an
ELISA plate assay.
Is My Antibody Specific?
Once an antibody has been purified, it is imperative to fully character-
ize its specificity. There are two common approaches that can be taken
to achieve this: (i) an ELISA assay or (ii) a western blot approach can
be performed in the presence or absence of varying concentrations of
different peptides. The first decision here concerns the source of substrate.
In an ELISA assay the wells are usually coated with the specific peptide
or protein that the antibody was raised against. In a western blot, the
blotted protein is typically purified histones from calf thymus (Sigma)
since these are a good source of many different methylated sites within
the histone N-terminal tails. In both the ELISA and western blot
approaches, competition peptides should be included to test the specificity
of the antibody. These competition peptides should include as wide a
range as possible of similar, but not identical, epitopes. For instance, if
the antibody to be characterized was raised against a H3 peptide (amino
acids 1–8) tri-methylated at lysine 4, the competition peptides should
include a non-methylated H3 peptide (1–8) as well as the mono-, di-, and
tri-methylated derivatives. Only the tri-methylated peptide should
block the binding of the antibody to its substrate. An example of a typical
result from a western blot approach is shown in Fig. 3. In addition to
these peptides, a number of other tri-methylated histone peptides should
also be tested to determine the absolute specificity of the antibody.
Fig. 3. Determining antibody specificity using peptide competition. Claf thymus histones
were loaded into a single wide well of a 20% SDS-PAG and resolved. They were then western
blotted to nitrocellulose. Strips were then cut from the blot and separately probed with an
anti-tri-methyl K4 of H3 antibody (available from www.abcam.com) in the presence (1 g/ml)
or absence of various peptides as indicated.
280 immunochemical assays of chromatin functions [18]
Histones that have been expressed in and purified from bacteria also
serve as excellent controls in assays designed to test the specificity of
anti-methylated histone antibodies. Because bacteria lack the methyltrans-
ferases that modify histones, the histones are unmethylated and therefore
should not be recognized by the antibodies.
Western Blots
An obvious application for the use of anti-methylated histone
antibodies is the characterization of substrates by western blot analysis.
The substrate protein may include (i) histones extracted from various
cellular sources or (ii) the methylated histone product of an in vitro
methylation assay (Fig. 4). Our protocol for this procedure is described
earlier.
Immunofluorescence
An investigator often wishes to ask whether a particular histone
methylation mark is enriched in specific structures within the genome, in
specific cells, or at specific points of the cell-cycle. One way to address
these issues is by using immunofluorescence. The main decisions here
regard which fixation and permeabilization techniques to use. The fol-
lowing is our general protocol that should serve as a reasonable starting
point.
1. Cells are grown under the appropriate condition in 25-cm2 slide
flasks (NUNC).
2. Aspirate growth medium and wash the cells twice with PBS.
3. The cells are then fixed in 2 ml of 4% (v/v) formaldehyde in PBS
for 20 min at room temperature. Alternatively, paraformaldeyde
may be used if formaldehyde fails to give reasonable results.
4. Wash cells three times (5 min each) with PBS.
5. Block/permeabilize the cells with 3% (w/v) BSA, 0.6% (v/v) Triton
X-100 in PBS for 15 min at room temperature. An alternative
detergent to consider is 0.2% (v/v) Tween 20, though we find that
this often fails to fully permeabilize the cells.
6. Wash three times (5 min each) in PBS.
7. Incubate with secondary antibody (conjugated to an appropriate
flour) in 3% (w/v) BSA in PBS for 1 h in the dark.
8. Wash once in PBS.
9. Hoechst 33258 in PBS is added to a final concentration of 1 g/ml
and incubated for 5 min at room temperature in the dark.
10. The cells are finally washed four times in PBS, each for 5 min.
11. The flask is then opened using the tool supplied by the
manufacturer and the resulting slide is mounted using standard
techniques.
282 immunochemical assays of chromatin functions [18]
19
S. J. Nielsen, R. Schneider, U. M. Bauer, A. J. Bannister, A. Morrison, D. O’Carroll,
R. Firestein, M. Cleary, T. Jenuwein, R. E. Herrera, and T. Kouzarides, Nature (London)
412, 561 (2001).
284 immunochemical assays of chromatin functions [18]
5. Wash the pellet once with 1 ml final wash buffer (20 mM Tris-HCl,
pH 8.0, 500 mM NaCl, 2 mM EDTA, 1% [v/v] Triton X-100, 0.1%
[w/v] SDS). Spin as earlier.
6. Elute the bound DNA by adding 450 l of freshly made elution
buffer (1% [w/v] SDS, 100 mM NaHCO3) and rotating on a wheel
for 15 min at room temperature. Spin down as earlier and remove
the supernatant into a fresh tube.
7. Add 5 l proteinase K (20 mg/ml) and heat with shaking at 65
for 4–5 h to reverse cross-linking (can freeze samples here if
necessary).
8. The samples are then phenol:chloroform extracted with an equal
volume of 1:1 phenol:choroform. Ten microliters of glycogen (5 mg/
ml stock solution) are added to aid precipitation and then 1/10th
volume of sodium acetate and 2 volumes of ethanol are added. The
precipitations are incubated on dry ice for 10 min and the DNA is
then pelleted by spinning at 13,000g for 10 min. The pellet is
carefully washed in 70% (v/v) ethanol and then dried. Finally the
DNA pellet is dissolved in 100 l TE (10 mM Tris-HCl) pH 8.0,
1 mM EDTA, and stored at 4 or 20 .
20
S. Rea, F. Eisenhaber, D. O’Carroll, B. D. Strahl, Z. W. Sun, M. Schmid, S. Opravil,
K. Mechtler, C. P. Ponting, C. D. Allis, and T. Jenuwein, Nature (London) 406, 593 (2000).
288 immunochemical assays of chromatin functions [18]
Acknowledgments
We would like to thank Darren Harper, Abcam Ltd. (www.abcam.com) for suggestions
and helpful advice relating to antibody production and ELISA assays, Michael Hendzel
(Alberta University, Canada) for supplying immunofluorescence images, and Graham
Bloomberg for advice on peptide synthesis. All our methylated peptides are synthesized by
Graham Bloomberg ([email protected]).
[19] acetylation microarray 289
1
N. Suka, Y. Suka, A. A. Carmen, J. Wu, and M. Grunstein, Mol. Cell 8, 473 (2001).
2
J. Wu, N. Suka, M. Carlson, and M. Grunstein, Mol. Cell 7, 117 (2001).
3
A. Wang, S. K. Kurdistani, and M. Grunstein, Science 298, 1412 (2002).
4
D. Kadosh and K. Struhl, Cell 89, 365 (1997).
5
S. E. Rundlett, A. A. Carmen, N. Suka, B. M. Turner, and M. Grunstein, Nature 392, 831
(998).
6
M. Vogelauer, J. Wu, N. Suka, and M. Grunstein, Nature 408, 495 (2000).
7
B. E. Bernstein, J. K. Tong, and S. L. Schreiber, Proc. Natl. Acad. Sci. USA 97, 13708
(2000).
8
T. R. Hughes, M. J. Marton, A. R. Jones, C. J. Roberts, R. Stoughton, C. D. Armour, H. A.
Bennett, E. Coffey, H. Dai, Y. D. He, M. J. Kidd, A. M. King, M. R. Meyer, D. Slade, P. Y.
Lum, S. B. Stepaniants, D. D. Shoemaker, D. Gachotte, K. Chakraburtty, J. Simon, M. Bard,
and S. H. Friend, Cell 102, 109 (2000).
9
S. K. Kurdistani, D. Robyr, S. Tavazoie, and M. Grunstein, Nat. Genet. 31, 248 (2002).
10
D. Robyr, Y. Suka, I. Xenarios, S. K. Kurdistani, A. Wang, N. Suka, and M. Grunstein, Cell
109, 437 (2002).
11
A. Hecht, S. Strahl-Bolsinger, and M. Grunstein, Nature 383, 92 (1996).
12
V. R. Iyer, C. E. Horak, C. S. Scafe, D. Botstein, M. Snyder, and P. O. Brown, Nature 409,
533 (2001).
13
B. Ren, F. Robert, J. J. Wyrick, O. Aparicio, E. G. Jennings, I. Simon, J. Zeitlinger,
J. Schreiber, N. Hannett, E. Kanin, T. L. Volkert, C. J. Wilson, S. P. Bell, and R. A. Young,
Science 290, 2306 (2000).
[19] acetylation microarray 291
14
A. Hecht, S. Strahl-Bolsinger, and M. Grunstein, Methods Mol. Biol. 119, 469 (1999).
[19] acetylation microarray 293
5. Immunoprecipitate acetylated chromatin fragments overnight at 4
on a nutator incubating 50 l of cell lysate with 2–5 l of a given
antiserum (see later). Then, add a suspension of 25 l of 50% (v/v)
protein A sepharose CL-4B beads (Amersham-Pharmacia),
equilibrated in lysis buffer and incubate an additional 2 h.
6. Pellet the sepharose beads for 30 s at room temperature in an
Eppendorf centrifuge (model 5415 C) at 735g (3000 rpm). Discard
the supernatant and wash successively for 5 min the beads on a
nutator with 500 l of the following solutions: twice in lysis buffer,
once in deoxycholate buffer (10 mM Tris-HCl, pH 8.0, 0.25 M LiCl,
0.5% (v/v) NP-40, 0.5% (w/v) sodium deoxycholate and 1 mM
EDTA, pH 8.0), and once in TE, pH 8.0. Pellet the beads and
discard the supernatant between each washing step as indicated
earlier. The whole procedure is carried out at room temperature.
7. Elute the immunoprecipitated chromatin fragment from the beads
and reverse crosslink overnight at 65 with 50 l elution buffer
(50 mM Tris-HCl, pH 8.0, 10 mM EDTA, pH 8.0, and 1% (v/v) SDS).
8. Add 50 l of TE, pH 8.0, 20 g glycogen and treat with 20 g
proteinase K for 2–3 h at 55 . Finally, extract DNA with 1 volume
phenol/chloroform/isoamyl alcohol [25:24:1 (v/v)] and ethanol
precipitate. Resuspend purified DNA into 50 l TE, pH 8.0 and
store at 20 .
Troubleshooting
The antibody specificity is the most critical aspect for any chromatin im-
munoprecipitation. The polyclonal antibodies raised against individual
acetylated lysine residues were prepared in our laboratory and are avail-
able at Upstate (http://www.upstate.com). Their specificity was verified
by ELISA assay and tested by ChrIP against histones mutated for the acet-
ylatable lysine.1 The titration of the antibody amount required for an im-
munoprecipitation has to be determined experimentally for all antibodies.
Sonication conditions will determine the resolution of the chromatin
immunoprecipitation (the higher the average fragment size is, the lower
the resolution will be). Shearing efficiency will also depend on the sonicator
brand. A pilot experiment should be performed in order to find the appro-
priate sonication settings. After sonication add 80 l elution buffer (see
step 7 for recipe) to a 20-l aliquot from the cleared cell lysate (step 4)
and incubate the tube overnight at 65 . Add 100 l TE, pH 8.0, 40 g pro-
teinase K and incubate the tubes for 2–3 h at 55 as indicated earlier (step
8). Resuspend DNA pellet (Input DNA) in 20 l TE, pH 8.0, after DNA
extraction and precipitation. Treat an aliquot (10 l) with RNase A
294 immunochemical assays of chromatin functions [19]
(10 g) 30 min at 37 and analyze the average DNA fragment size on a
1.5% agarose gel. Alter accordingly the sonication settings if DNA frag-
ments are too large. Store the remaining RNase untreated DNA at 20
for later PCR amplification if needed (see later).
15
S. K. Kurdistani and M. Grunstein, Methods 31, 90 (2003).
[19] acetylation microarray 295
1. Use small 0.2-ml PCR tubes and perform all reactions in a PCR
thermocycler. Add 2 l 5 Sequenase buffer (200 mM Tris-HCl, pH
7.5, 100 mM MgCl2, and 250 mM NaCl) and 1 l (40 pmol)
oligonucleotide (50 -GTTTCCCAGTCACGATCNNNNNNNNN-
30 ) to 7 l immunoprecipitated DNA. Denature DNA for 2 min at
94 , cool the tubes down to 8 , and incubate at 8 for an additional
2 min. Pause the PCR machine and add 5 l of reaction mix (1
Sequenase buffer, 0.9 mM dNTPs, 15 mM DTT, 0.75 g BSA, and
4 U Sequenase 2.0). The oligonucleotide is annealed to DNA by
slowly raising the temperature from 8 to 37 over a period of 8 min.
DNA synthesis is allowed to proceed for an extra 8 min at 37 .
Repeat once the whole denaturation-annealing-elongation process.
However, since the DNA polymerase is heat sensitive, add 4 U
Sequenase after the denaturation step at 94 . Finally, stop the
reaction by diluting the reaction with 35 l TE, pH 8.0. DNA (step 1
DNA) can be stored at 20 or used immediately for the PCR
reaction described later.
2. The following step of the DNA amplification procedure is a regular
PCR reaction using the fixed linker oligonucleotide sequence (50 -
GTTTCCCAGTCACGATC-30 ) and DNA from step 1 as a
template. Transfer an aliquot of 15 l step 1 DNA into a fresh
0.5-ml PCR tube and carry the reaction in a final volume of 100 l
(20 mM Tris-HCl, pH 8.4, 50 mM KCl, 2 mM MgCl2, 1.25
nanomoles oligonucleotide, 0.25 mM dNTPs, and 5 U recombinant
Taq polymerase [Invitrogen]). Denature DNA at 92 for 30 s,
anneal with two consecutive short 30-s steps at 40 and 50 ,
respectively, and elongate at 72 during 90 s. Repeat this cycle 24
times and allow the final elongation to proceed for another 10 min
at 72 .
3. Purify the PCR reaction through columns (QIAquick PCR
purification kit, Qiagen) as indicated by the manufacturer and elute
DNA with 50 l 10 mM Tris-HCl, pH 8.5.
4. Estimate DNA yield under UV light by spotting serial dilutions of
DNA in a Petri dish (0.5 l DNA mixed with 0.5 l EtBr (10 g/ml)
along with a standard of known concentration (2-fold serial dilution
from 100 ng/l to 6.25 ng/l) and test the average size of DNA (5 l
aliquot) after PCR on a 1.5% (w/v) agarose gel. The probe average
size is reduced from 500 bp (after sonication) down to 300 bp. This
is due to the random incorporation of the oligonucleotide into DNA
during step 1.
[19] acetylation microarray 297
Troubleshooting
We routinely obtain between 3 and 5 g DNA after amplification by
PCR. It is not absolutely necessary to have such a DNA yield since the la-
beling reaction described later requires 500 ng DNA. However, if the
amount of DNA is not satisfactory, try to start the amplification procedure
with more material or scale-up the chromatin immunoprecipitation. We do
not advise increasing the number of PCR cycles since it is kept relatively
low in order to ensure amplification linearity.
The final PCR purification through columns (QIAquick PCR purifica-
tion kit, Qiagen) is critical since unincorporated oligonucleotides and
dNTPs may interfere with the subsequent labeling reaction efficiency.
For some microarray applications, it may be necessary to compare im-
munoprecipitated DNA with input DNA (i.e., study of a wild-type strain
acetylation profile instead of analyzing the changes of acetylation between
a WT and a HDAC mutant). The Input DNA is prepared from 20 l cleared
lysate as indicated earlier (see the Troubleshooting section for Chromatin
Immunoprecipitation (ChrIP or ChIP)). Dilute 150-fold an aliquot of the
Input DNA prior to starting the amplification procedure (step 1).
Troubleshooting
In our hands, the labeling procedure is reliable and successful most of
the time. The color of the purified probe is the best indicator for the pro-
cedure success. The combination of the two probes should give a dark
purple color after concentration. Do not discard the flow-through if the
probe is colorless after the first 7 min centrifugation step through the
microcon-30 filter, as it is possible that the filter is defective. Simply reload
the flow-through onto another filter and centrifuge again. If the problem
persists, this is a clear indication that the labeling was not efficient and
the probe should not be used for hybridization. Similarly, a pink (Cy3-
labeling) or a blue (Cy5-labeling) color indicates that one of the probes
was not labeled properly. Change the vial of dye if you suspect the lack
of efficient labeling might stem from a bad batch. The dNTP mix is also
an important factor for labeling. We have consistently observed a reduction
in its efficiency when using the same dNTP mix stock several times. We
strongly advise to prepare it fresh shortly before the labeling reaction.
We have observed that pre-hybridization of the microarray slide is not
required when used for acetylation microarrays. Some antibodies used for
immunoprecipitation of epitope-tagged protein (i.e., anti-hemagglutinin)
appear to create unacceptable hybridization background levels. The
following pre-hybridization conditions can be applied in order to alleviate
or reduce the background problem. Incubate the slides in 3.5 SSC, 0.1%
(v/v) SDS, and 10 mg/ml BSA in a staining jar for 20 min at 44 . Rinse
the slides briefly first with water then with isopropanol. Dry the slides by
centrifugation on a microtiter plate carrier as indicated earlier (step 4).
18
N. Suka, K. Luo, and M. Grunstein, Nat. Genet. 32, 378 (2002).
[19] acetylation microarray 301
Fig. 2. Histone acetylation sites correlate differently with transcription resulting from
rpd3. This figure illustrates the scaling of different acetylation data sets using percentile
ranking in order to compare them with a transcription data set. Due to inherent noise in
microarray data, absolute correlation analysis between acetylation and transcription is not
very informative. A moving average analysis that greatly reduces noise can be used to extract
general trends. The moving average (window size, 100 data point; step, 1 data point)
percentile rank of acetylation enrichment is plotted as a function of transcription increase
resulting from rpd3 (B. E. Bernstein, J. K. Tong, and S. L. Schreiber, Proc. Natl. Acad. Sci.
USA 97, 13708 (2000)). Acetylation data are plotted for H4 K5 (dark blue), H4 K12 (red), H4
K16 (orange), and H3 K18 (light blue). Control corresponds to a comparison of two sets of
probes amplified from the immunoprecipitation of acetylated H4 K12 in the WT strain and
labeled separately with Cy3 and Cy5 prior to hybridization. Data show that increased
acetylation at histone H4 K5 and K12 is associated most directly with increased gene
transcription in rpd3. H4 K16 show the poorest correlation with gene activity (RPD3
disruption has no significant effects on the status of H4 K16 acetylation) (D. Robyr, Y. Suka,
I. Xenarios, S. K. Kurdistani, A. Wang, N. Suka, and M. Grunstein, Cell 109, 437 (2002)).
Reprinted with permission from D. Robyr et al., Cell 109, 437–446 (2002).
ORF it regulates. Some intergenic regions are located between two diver-
gent ORFs and are by default assigned to both ORFs. Other intergenic
regions are located more than 1 kb away from any ORF. These are
considered orphans and are not assigned to any particular ORF.
Concluding Remarks
Acetylation microarrays have proved to be extremely useful in dis-
covering new functions for the yeast HDACs, functions that would have
been much more difficult to identify using classical genetic or molecular
biology techniques.10 Although the approach described here focuses on
acetylation, it can be extended to other histone modifications as well. The
19
M. B. Eisen and P. O. Brown, Methods Enzymol. 303, 179 (1999).
[19] acetylation microarray 303
Fig. 3. Binding arrays are complementary to acetylation arrays. The moving average
(window size, 100 data point; step, 1 data point) of Rpd3 enrichment (binding) over intergenic
regions is plotted as a function of percentile rank of H3 K18 acetylation in rpd3. Data sets
with (+RP) and without (RP) ribosomal protein genes are plotted as indicated. Rpd3 binds
strongly at the promoter of ribosomal protein genes in logarithmically growing cells where
these genes are highly active but under the same conditions, has little or no effect on
acetylation of these promoters (S. K. Kurdistani, D. Robyr, S. Tavazoie, and M. Grunstein,
Nat. Genet. 31, 248 (2002); D. Robyr, Y. Suka, I. Xenarios, S. K. Kurdistani, A. Wang, N. Suka,
and M. Grunstein, Cell 109, 437 (2002)) or the expression of the RP genes (B. E. Bernstein,
J. K. Tong, and S. L. Schreiber, Proc. Natl. Acad. Sci. USA 97, 13708 (2000)). Thus, the
ribosomal protein gene promoters as targets of Rpd3 are only detectable by binding arrays.
This clearly illustrates the importance of combining different types of arrays (binding,
acetylation, and expression) to fully comprehend HDAC function. Reprinted with permission
from S. K. Kurdistani et al., Nat. Gen. 31, 248–254 (2002).
20
L. Zeng and M. M. Zhou, FEBS Lett. 513, 124 (2002).
21
T. R. Hughes, C. J. Roberts, H. Dai, A. R. Jones, M. R. Meyer, D. Slade, J. Buchard,
S. Dow, T. R. Ward, M. J. Kidd, S. H. Friend, and M. J. Marton, Nat. Genet. 25, 333 (2000).
20
L. Zeng and M. M. Zhou, FEBS Lett. 513, 124 (2002).
21
T. R. Hughes, C. J. Roberts, H. Dai, A. R. Jones, M. R. Meyer, D. Slade, J. Buchard,
S. Dow, T. R. Ward, M. J. Kidd, S. H. Friend, and M. J. Marton, Nat. Genet. 25, 333 (2000).
genes with DNA microarray technology. This method will enable research-
ers to overcome limitations in the study of eukaryotic transcriptional regu-
lation by allowing the identification of direct, physiological targets of
DNA-binding proteins in an unbiased, genome-wide manner.
Procedures
Overview
Genome-wide location analysis, described in detail in this chapter, is
based on the rapid purification of specific genomic fragments associated
with a particular protein followed by detection of enriched sequences with
DNA microarray technology (see Fig. 1). Intact, living cells are treated
with formaldehyde, a cross-linking agent that results in the covalent and
reversible linkage between genomic DNA and associated proteins.1 The
cross-linked chromatin is extracted after disruption of cell and nuclear
membranes and fragmented into an average length of 500 bp by sonication.
The chromatin associated with a factor of interest is then purified through
immunoprecipitation using a specific antibody against the protein. Subse-
quently, the cross-links are reversed, and the genomic DNA in the resulting
sample is purified by successive steps to remove protein and RNA. DNA
fragments are blunt-ended by T4 DNA polymerase to allow ligation of a
universal oligonucleotide linker, and all DNA fragments in the sample
are amplified through polymerase chain reaction (PCR) in an unbiased
fashion. As a control, an aliquot of the input chromatin DNA is processed
1
V. Orlando, Trends Biochem. Sci. 25, 99 (2000).
306 immunochemical assays of chromatin functions [20]
2
B. Ren, H. Cam, Y. Takahashi, T. Volkert, J. Terragni, R. A. Young, and B. D. Dynlacht,
Genes Dev. 16, 245 (2002).
[20] use of ChIP in location analysis 307
Solutions
Cross-linking solution: 11% formaldehyde, 0.1 M
NaCl, 1 mM Na-EDTA, 0.5 mM
Na-EGTA, 50 mM HEPES,
pH 8.0
Lysis buffer 1: 0.05 M HEPES-KOH, pH 7.5,
0.14 M NaCl, 1 mM EDTA,
10% glycerol, 0.5% NP-40,
0.25%, Triton X-100, with
protease inhibitor cocktail
(Roche Applied Science)
Lysis buffer 2: 0.2 M NaCl, 1 mM EDTA,
0.5 mM EGTA, 10 mM Tris, pH
8, protease inhibitor cocktail
Lysis buffer 3: 1 mM EDTA, 0.5 mM EGTA,
10 mM Tris-HCl, pH 8, protease
inhibitor cocktail
Proteinase K stock solution: 20 mg/ml proteinase K (Sigma)
50 mM Tris-HCl, pH 8.0, 1.5 mM
calcium acetate
RIPA buffer: 50 mM HEPES, pH 7.6, 1 mM
EDTA, 0.7% DOC, 1% NP-40,
0.5 M LiCl, protease inhibitor
cocktail
Elution buffer: 50 mM Tris, pH 8, 10 mM EDTA,
1% SDS
Hybridization buffer 1: 2.2 SSC, 0.22% SDS
Hybridization buffer 2: 70% formamide, 3 SSC, 14.3%
dextran sulfate
308 immunochemical assays of chromatin functions [20]
Immunoprecipitation of Chromatin
Magnetic beads (Dynal) pre-bound to polyclonal antibodies are used to
immunoprecipitate the DNA associated with the protein of interest. To
prepare the magnetic beads, 100 l of sheep anti-rabbit IgG-conjugated
Dynabeads (Dynal) are first washed three times with cold PBS containing
5 mg/ml bovine serum albumin (BSA) and then re-suspended in 5 ml of
cold PBS. Ten micrograms of rabbit polyclonal antibody are added to the
mixture and incubated overnight on a rotating platform at 4 . After collect-
ing the magnetic beads by centrifugation and washing three times with cold
PBS containing 5 mg/ml BSA, the beads are re-suspended in 100 l of
cold PBS with 5 mg/ml BSA and are then ready for immunoprecipitation.
In an Eppendorf tube, 2 mg of soluble chromatin are first adjusted to
0.1% Triton X-100, 0.1% sodium deoxycholate, and 1 mM PMSF, then
mixed with 100 l of magnetic beads pre-bound to the antibody. The mix-
ture is incubated at 4 overnight on a rotating platform. The magnetic
beads are then collected using a magnet (Dynal), and the supernatant is
removed by aspiration. To remove material non-specifically bound to the
beads, 1 ml of RIPA buffer is added to the tube, and the beads are gently
re-suspended on a rotating platform in the cold room. The magnetic beads
are again collected with the magnet and washed with RIPA buffer a total of
five times. After washing once with 1 ml of TE, the beads are collected by
centrifugation at 2000g for 3 min and re-suspended in 50 l of elution
buffer. To elute precipitated chromatin from the beads, the tubes are incu-
bated at 65 for 10 min with constant agitation then centrifuged for 30 s at
2000g. Forty microliters of supernatant are removed and mixed with 120 l
of TE with 1% SDS. This solution is incubated at 65 overnight to reverse
the cross-links. As a chromatin input control, 100 g of chromatin are
mixed with 120 l of TE containing 1% SDS in a separate tube and
incubated at 65 overnight.
Purification of DNA
After reversal of cross-links, proteins in the DNA sample are removed
by incubation with 120 l of Proteinase K solution (2% glycogen, 5%
Proteinase K stock solution, and TE) for 2 h at 37 . The sample is then ex-
tracted twice with phenol and once with 24:1 chloroform/isoamyl alcohol.
The sample is adjusted to 200 mM NaCl. After ethanol precipitation,
the DNA is dissolved in 30 l of TE containing 10 g of DNase-free
RNase A and incubated for 2 h at 37 . The DNA at this step can be further
purified with a Qiagen PCR kit (Qiagen).
310 immunochemical assays of chromatin functions [20]
3
C. J. Roberts, B. Nelson, M. J. Marton, R. Stoughton, M. R. Meyer, H. A. Bennett, Y. D.
He, H. Dai, W. L. Walker, T. R. Hughes et al., Science 287, 873 (2000).
314 immunochemical assays of chromatin functions [20]
where a1,2 are the intensities measured in the two channels for each
spot and Xnorm is the normalized X for each spot. The weights for each spot
are then defined as
wi ¼ 1=2i (4)
The averaged log(ratio) is then calculated using the following formula
P
i¼1;n wi xi
x¼ P ; (5)
i¼1;n wi
where n is the total number of experiments, xi and wi are the log(ratio) and
weight, respectively, for a particular spot in each experiment; x is normally
distributed and the averaged P value is then calculated using Eq. (2), where
the variable for Erf( ) function is x instead. Target genes were selected on
the basis of significant P values (e.g., less than 0.002) and binding ratios
(e.g., greater than 2) in replicate experiments.
Conclusion
In summary, the method that we describe in this chapter is a direct,
high-throughput and general approach to locate protein-DNA interactions
in mammalian cells. Since most transcription factors function by binding
to DNA and regulating gene expression, this method will allow for the
rapid, specific identification of physiologically relevant target genes and
provide key insights into the molecular mechanisms through which they
function. We expect that this technique will have widespread applications
in deciphering global gene regulatory networks in mammalian cells as it
has in yeast.4 In addition to revealing the location of promoter and en-
hancer binding proteins, the method can also be applied to examine regu-
latory elements bound by the DNA replication and recombination
machinery and nucleosome modification complexes, and it should also
allow a systematic identification of sites of histone modification and
DNA methylation.5–9
4
T. I. Lee, N. J. Rinaldi, F. Robert, D. T. Odom, Z. Bar-Joseph, G. K. Gerber, N. M.
Hannett, C. T. Harbison, C. M. Thompson, I. Simon, J. Zeitlinger, E. G. Jennings, H. L.
Murray, D. B. Gordon, B. Ren, J. J. Wyrick, J. B. Tagne, T. L. Volkert, E. Fraenkel, D. K.
Gifford, and R. A. Young, Science 298, 799 (2002).
5
B. E. Bernstein, E. L. Humphrey, R. L. Erlich, R. Schneider, P. Bouman, J. S. Liu,
T. Kouzarides, and S. L. Schreiber, Proc. Natl. Acad. Sci. USA 99, 8695 (2002).
6
Y. Blat and N. Kleckner, Cell 98, 249 (1999).
7
J. L. Gerton, J. DeRisi, R. Shroff, M. Lichten, P. O. Brown, and T. D. Petes, Proc. Natl.
Acad. Sci. USA 97, 11383 (2000).
[21] screening of chromatin immunoprecipitates 315
Acknowledgments
We thank H. Cam, Y. Takahashi, T. Volkert, and members of Richard Young’s
laboratory for help with the development of techniques described in this chapter. Work in
BDD’s laboratory was supported by the American Cancer Society.
8
D. Robyr, Y. Suka, I. Xenarios, S. K. Kurdistani, A. Wang, N. Suka, and M. Grunstein, Cell
109, 437 (2002).
9
J. J. Wyrick, J. G. Aparicio, T. Chen, J. D. Barnett, E. G. Jennings, R. A. Young, S. P. Bell,
and O. M. Aparicio, Science 294, 2357 (2001).
1
R. V. Davuluri, I. Grosse, and M. Q. Zhang, Nat. Genet. 29, 412 (2001).
2
A. E. Kel, O. V. Kel-Margoulis, P. J. Farnham, S. M. Bartley, E. Wingender, and M. Q.
Zhang, J. Mol. Biol. 309, 99 (2001).
3
S. Aerts, G. Thijs, B. Coessens, M. Staes, Y. Moreau, and B. De Moor, Nucleic Acids Res.
31, 1753 (2003).
4
P. C. Fernandez, S. R. Frank, L. Wang, M. Schroeder, S. Liu, J. Greene, A. Cocito, and
B. Amati, Genes Dev. 17, 1115 (2003).
Acknowledgments
We thank H. Cam, Y. Takahashi, T. Volkert, and members of Richard Young’s
laboratory for help with the development of techniques described in this chapter. Work in
BDD’s laboratory was supported by the American Cancer Society.
8
D. Robyr, Y. Suka, I. Xenarios, S. K. Kurdistani, A. Wang, N. Suka, and M. Grunstein, Cell
109, 437 (2002).
9
J. J. Wyrick, J. G. Aparicio, T. Chen, J. D. Barnett, E. G. Jennings, R. A. Young, S. P. Bell,
and O. M. Aparicio, Science 294, 2357 (2001).
1
R. V. Davuluri, I. Grosse, and M. Q. Zhang, Nat. Genet. 29, 412 (2001).
2
A. E. Kel, O. V. Kel-Margoulis, P. J. Farnham, S. M. Bartley, E. Wingender, and M. Q.
Zhang, J. Mol. Biol. 309, 99 (2001).
3
S. Aerts, G. Thijs, B. Coessens, M. Staes, Y. Moreau, and B. De Moor, Nucleic Acids Res.
31, 1753 (2003).
4
P. C. Fernandez, S. R. Frank, L. Wang, M. Schroeder, S. Liu, J. Greene, A. Cocito, and
B. Amati, Genes Dev. 17, 1115 (2003).
11
V. R. Iyer, C. E. Horak, C. S. Scafe, D. Botstein, M. Snyder, and P. O. Brown, Nature 409,
533 (2001).
12
J. D. Lieb, X. Liu, D. Botstein, and P. O. Brown, Nat. Genet. 28, 327 (2001).
13
B. Ren, F. Robert, J. J. Wyrick, O. Aparicio, E. G. Jennings, I. Simon, J. Zeitlinger,
J. Schreiber, N. Hannett, E. Kanin, T. L. Volkert, C. J. Wilson, S. P. Bell, and R. A. Young,
Science 290, 2306 (2000).
14
H. H. Ng, F. Robert, R. A. Young, and K. Struhl, Genes Dev. 16, 806 (2002).
15
M. Damelin, I. Simon, T. I. Moy, B. Wilson, S. Komili, P. Tempst, F. P. Roth, R. A. Young,
B. R. Cairns, and P. A. Silver, Mol. Cell. 9, 563 (2002).
16
J. J. Wyrick, J. G. Aparicio, T. Chen, J. D. Barnett, E. G. Jennings, R. A. Young, S. P. Bell,
and O. M. Aparicio, Science 294, 2357 (2001).
17
T. I. Lee, N. J. Rinaldi, F. Robert, D. T. Odom, Z. Bar-Joseph, G. K. Gerber, N. M.
Hannett, C. T. Harbison, C. M. Thompson, I. Simon, J. Zeitlinger, E. G. Jennings, H. L.
Murray, D. B. Gordon, B. Ren, J. J. Wyrick, J. Tagne, T. L. Volkert, E. Fraenkel, D. K.
Gifford, and R. A. Young, Science 298, 799 (2002).
18
P. Kapranov, S. E. Cawley, J. Drenkow, S. Bekiranov, R. L. Strausberg, S. P. Fodor, and
T. R. Gingeras, Science 296, 916 (2002).
19
J. L. Rinn, G. Euskirchen, P. Bertone, R. Martone, N. M. Luscombe, S. Hartman, P. M.
Harrison, F. K. Nelson, P. Miller, M. Gerstein, S. Weissman, and M. Snyder, Genes Dev. 17,
529 (2003).
318 immunochemical assays of chromatin functions [21]
Chromatin Immunoprecipitation
We have recently published a detailed protocol that we have found
will efficiently and specifically immunoprecipitate sequences associated
20
B. Ren, H. Cam, Y. Takahashi, T. Volkert, J. Terragni, R. A. Young, and B. D. Dynlacht,
Genes Dev. 16, 245 (2002).
21
A. S. Weinmann, P. S. Yan, M. J. Oberley, T. H. Huang, and P. J. Farnham, Genes Dev. 16,
235 (2002).
22
J. Wells, P. S. Yan, M. Cechvala, T. Huang, and P. J. Farnham, Oncogene 22, 1445 (2003).
23
D. Y. L. Mao, J. Watson, P. S. Yan, D. Barsyte-Lovejoy, F. Khosravi, W. W. Wong,
P. Farnham, T. H. Huang, and L. Z. Penn, Curr. Biol. 13, 882 (2003).
[21] screening of chromatin immunoprecipitates 319
Fig. 1. Preparation of samples for use in a ChIP-CpG assay (Steps I–III). Step I: cells
are crosslinked in culture (A), which creates protein-protein and protein-DNA crosslinks.
Nuclei are then isolated (B) and sonicated to create chromatin fragments of approximately
500–1000 bp. The chromatin is then immunoprecipitated with antibodies specific to the
protein of interest; an IgG control sample can also be prepared (C). After extensive washing,
the crosslinks are reversed and the DNA is purified, resulting in pools of sequences that are
enriched by virtue of their binding to the protein of interest. Chromatin is also purified from
crosslinked DNA that has not been immunoprecipitated; this serves as an input DNA control
(D). At this point, PCR reactions should be performed using a positive control (see Fig. 3A).
Step II: the pools of DNA are separately subjected to ligation-mediated PCR to generate
amplicons (E). At this point, PCR reactions should be performed using the same set of
primers as above (see Fig. 3B). Step III: the IP amplicon is labeled with Cy5, and an equal
amount of the total input amplicon is labeled with Cy3; the two will be combined to hybridize
to a CpG island array (see Fig. 2). The IgG amplicon can also be labeled with Cy5 and
hybridized with the input amplicon labeled with Cy3 as an additional control (see text).
320 immunochemical assays of chromatin functions [21]
Fig. 2. Preparation and hybridization of CpG arrays (Steps IV–VI). Step IV: CpG islands
are spotted onto glass slides and post-processed according to the manufacturer’s protocols.
Step V: the Cy5- and Cy3-labeled samples are hybridized to the CpG-island microarray in the
presence of CoT-1 DNA overnight at 60 . The slides are then washed to remove unbound
labeled sample and then dried. Step VI: the fluorescent intensity of each feature is read with
an appropriate scanner and data are imported into analysis program for further data analysis.
Features with intensities significantly higher in the ChIP-labeled samples relative to the input
sample are regarded as putative binding targets. However, if a feature also is positive on the
IgG versus input control array, it is removed from the list of targets.
24
M. J. Oberley and P. J. Farnham, Meth. Enzymol. 371, 579 (2003).
25
A. S. Weinmann and P. J. Farnham, Methods 26, 37 (2002).
[21] screening of chromatin immunoprecipitates 321
Fig. 3. ChIP controls performed prior to hybridization. (A) A traditional ChIP assay should
be performed prior to subjecting the samples to LMPCR. In this example, ChIP was performed
on 293 cells according to published protocols25 using antibodies specific for E2F6 and RNA
polymerase II (Pol2); a control was also performed using only a secondary antibody (IgG).
After part D of Step I (see Fig. 1), PCR was performed with primers specific to the UXT or
MYC promoters and the resulting signal was plotted as a fraction of the signal obtained using
0.2% of the total input chromatin. As expected, E2F6 is highly enriched on the UXT promoter
but not the MYC promoter. Pol II occupancy is robustly detected on both promoters, in
accordance with the fact that both genes are expressed in 293 cells (data not shown). Neither
promoter is enriched by non-specific IgG alone. Therefore, these ChIP samples are
appropriate for use in subsequent LMPCR reactions. (B) A PCR reaction should be
performed after the samples are subjected to LMPCR. In this case, LMPCR was performed as
described in section (Chromatin Amplicon Labeling) on the samples used in panel (A). Ten
nanograms of each amplicon (E2F6, Pol II, IgG, and total) was then subjected to PCR with
primers specific for the UXT and MYC promoters. The enrichment of UXT sequences seen in
panel (A) by the E2F6 antibody is maintained after LMPCR, as the signal seen is at least 3-fold
higher than the starting input. The E2F6 signal is not enriched relative to the input amplicon on
the MYC promoter. These results indicate the LMPCR procedure did not introduce significant
bias into the system and that the samples are appropriate for hybridization to an array.
one ChIP assay can be efficiently amplified to give amplicons that accur-
ately represent the starting population (see Fig. 3B). The ligation-mediated
PCR technique simply involves blunt ending the chromatin, ligating a uni-
directional double-stranded oligonucleotide linker, and PCR amplifying
the resultant DNA population. Other groups have utilized random priming
of the starting chromatin material with comparable results.12 Recently, a
promising amplification method has been developed that utilizes in vitro
transcription of the chromatin templates to create RNA amplicons.26
Because this procedure results in a linear rather than exponential amplifi-
cation, this method has the potential to reduce potential bias introduced
by PCR-based amplification schemes. Indeed, this form of sample amplifi-
cation is commonly used in microarray studies of gene expression where
mRNA samples are limited.27
To ensure that the LMPCR reaction has not resulted in a loss of differ-
ence between the experimental and control samples, we amplify the IP
samples as well as a small portion (10–100 ng) of the starting input DNA.
This amplified input DNA will be used as a reference on the CpG microar-
ray. We also commonly amplify a non-specific IgG ChIP control reaction
to demonstrate that enrichment is due to the specific antibody-epitope
interaction. Therefore, three different samples can be prepared that can
be used in two different hybridizations comparisons (i.e., IP versus input
and IgG versus input). Please note that this LMPCR amplification step is
a new addition to our protocol; to call attention to this fact, both Day 1
and Day 2 have been identified with an exclamation point; all steps in these
sections are new.
! Day 1:
1. The two unidirectional linkers oligoJW102 (50 gcggtgacccgggagatct-
gaattc 30 ) and oligoJW103 (50 gaattcagatc 30 ) are annealed by combining
6.7 l of 100 M oligoJW102, 6.7 l of 100 M oligoJW103, and 86.6 l
H2O. This mixture is boiled for 5 min in a water bath, then allowed to
slowly cool to room temperature. The annealed linker can be stored
indefinitely at 20 .
2. Because sonication of chromatin creates overhanging ends, the
chromatin is blunt ended by mixing the chromatin sample derived from an
experimental sample, an IgG sample, or 10 ng of input with the following:
11 l T4 DNA polymerase buffer, 5 l 10 BSA, 5 l 2 mM dNTPs, 1 l
T4 DNA polymerase (New England Biolabs, M0203S), and H2O to 110 l.
26
B. Bernstein and S. S. Schreiber, Meth. Ezymol. 376, 350 (2004).
27
C. C. Xiang, M. Chen, O. A. Kozhich, Q. N. Phan, J. M. Inman, Y. Chen, and M. J.
Brownstein, Biotechniques 34, 386 (2003).
324 immunochemical assays of chromatin functions [21]
This mixture is placed at 37 for 45–60 min, and then each sample is
purified with a Qiaquick PCR purification kit, according to manufacturers
protocol, eluting with 30 l of elution buffer.
3. Next, double-stranded unidirectional linker from Step 1 is ligated to
the blunted chromatin by adding 27 l of the blunted chromatin, 10.3 l
H2O, 5 l 10 T4 DNA ligase buffer, 6.7 l annealed JW102/JW103
linker, and 1.0 l T4 DNA ligase (New England Biolabs, M0202S). This
mixture is placed at 16 overnight.
! Day 2:
4. The following morning, each sample is purified with the Qiaquick
PCR purification kit, as earlier, eluting with 30 l of elution buffer.
5. The linker-ligated chromatin is then PCR amplified to create stocks
of amplicons from which virtually unlimited amounts of DNA can be
generated for use with the CpG-island microarrays. Each PCR reaction
contains 5 l 10 Taq polymerase buffer, 7 l 2 mM dNTPs, 3 l MgCl2,
6.5 l betaine, 2.5 l oligoJW102 (20 M), 1 l Taq (Promega, M1861),
and 25 l of the linker-ligated chromatin.
6. PCR is performed using the following protocol:
55 for 2 min #
28
I. P. Ioshikhes and M. Q. Zhang, Nat. Genet. 26, 61 (2000).
29
S. H. Cross, J. A. Charlton, X. Nan, and A. P. Bird, Nat. Genet. 6, 236 (1994).
30
P. S. Yan, C. M. Chen, H. Shi, F. Rahmatpanah, S. H. Wei, and T. H. Huang, J. Nutr. 132,
2430S (2002).
[21] screening of chromatin immunoprecipitates 329
Day 2:
11. To remove the coverslip, the microarray(s) are inverted in a glass
dish filled with 1 SSC and 0.1% SDS preheated to 50 . If no drying has
occurred overnight, the Lifterslip will immediately slide off.
12. The arrays are then agitated in 1 SSC, 0.1% SDS at 50 for 5 min,
followed by agitation in 1 SSC, 0.1% SDS at room temperature for
5 min. Finally, the arrays are washed with 0.2 SSC for 5 min at room
temperature to remove residual SDS (which autofluoresces), and then
dried by centrifugation at 600 rpm for 5 min in a 50-ml conical tube.
13. The arrays are scanned immediately using an Axon 4000B scanner
(Axon Instruments).
Analysis
The identification of positive clones is a multi-step procedure, first re-
quiring normalization of the Cy5 and Cy3 channels on the array because
of potential labeling efficiency differences, followed by clone selection,
confirmation, and ultimately, functional analysis.
1. The hybridized microarrays are analyzed using the Genepix Pro 4.0
(Axon Instruments) software package. This provides a set of raw values for
each feature on each array. To identify clones that are selectively enriched
during IP relative to the starting population, one must first normalize the
Cy5 and Cy3 channels on the array. To do so, we make the assumption that
the vast majority of loci will not be enriched in the IP relative to total; thus,
we can normalize across the entire array, by taking the ratio of the medians
for all features and normalizing them to unity.
2. After normalization, a ratio is generated that is the intensity in the
Cy5 channel minus background divided by the intensity in the Cy3 channel
minus background. Features in areas of the array that are obviously
blemished are manually flagged and excluded from the data set, as are
features with signal that have low intensities (<500). The ratio is generated
for all features that meet this criterion, and features with ratios above 2 are
isolated for further analysis.
3. Because of the complexity of the mammalian genome and the
relatively large number of steps in this procedure, false positives are
inevitable. Great care must be taken to reduce these errors to prevent
further wasting of time and resources following up on potential artifacts. It
is imperative to reduce the error rate by repeating the entire procedure at
least three times, starting with independent chromatin IPs from a different
batch of crosslinked cells each time (see Fig. 4B).
[21] screening of chromatin immunoprecipitates 331
This system allows one to perturb a normal biological system without some
of the concerns with traditional gene-ablation technology where the
knock-out of a specific gene may be compensated for during development
of the animal by upregulation of other family members. One could use
RNAi to specifically knock-down the expression of the DNA binding
factor of interest and then use RNase-protection, Northern, or traditional
microarray gene expression analysis to determine whether the lack of
binding to a putative target has a functional consequence. Shi et al.33 have
developed protocols that allow the CpG-island microarray to be probed
with mRNA that utilizes RNA ligase-mediated full-length cDNA synthesis.
With this protocol, it will be possible to examine the effect of DNA binding
factor knock-down on the resultant mRNA expression of the identified
target genes using the same CpG-island microarrays as used for the target
gene identification.
Conclusions
Studies to date employing the ChIP-chip approach using human cells
have been limited by the need to refine the methodology used for similar
studies of less complex eukaryotes. However, great strides have been made
recently in overcoming issues related to sample size and genomic complex-
ity. A major problem in studying human transcription factors is the fact
that no appropriate microarrays have been widely available that can be
used to study intergenic DNA; that is, all commercially available microar-
rays have allowed only the detection of transcribed sequences. Although
individual investigators have begun spotting promoter-specific13 or pro-
moter-enriched30 arrays, individual labs cannot serve as a source for arrays
for all the investigators who wish to employ the ChIP-chip approach. Un-
fortunately, commercial microarray companies have not yet entered into
the production of human promoter arrays. However, now that the feasibil-
ity of using such arrays has been demonstrated, greater efforts to provide
investigators with a comprehensive set of human intergenic DNA chips
must be made. The Encode (Encyclopedia of DNA Elements) project,
which has the long-term goal of identifying all functional elements in the
human genome, has chosen 30 Mb (approximately 1% of the human
genome) for preliminary analysis (see http://www.genome.gov/ Pages/
Research/ENCODE/). It can be hoped that the technologies and products
arising from such projects will allow much greater access of all investigators
to the required reagents needed for a thorough analysis of the regulatory
elements encoded in the human genome.
33
H. Shi, P. S. Yan, C. M. Chen, F. Rahmatpanah, C. Lofton-Day, C. W. Caldwell, and T. H.
Huang, Cancer Res. 62, 3214 (2002).
334 immunochemical assays of chromatin functions [22]
Acknowledgments
We would like to thank P. S. Yan and T. Huang for their generous gift of the CpG arrays.
M.J.O would like to thank A. Kirmizis for help adapting the LMPCR protocol and for stimulating
discussion. The Farnham lab would also like to thank A. Kirmizis and A. S. Weinmann for critical
readings of this manuscript. The Microarray Centre, which would like to thank J. Woodgett and
Neil Winegardin for technical input, was supported by ORDCF and Genome Canada. This work
was supported in part by Public Health Service grants CA22484, CA09135, and CA45240.
Acknowledgments
We would like to thank P. S. Yan and T. Huang for their generous gift of the CpG arrays.
M.J.O would like to thank A. Kirmizis for help adapting the LMPCR protocol and for stimulating
discussion. The Farnham lab would also like to thank A. Kirmizis and A. S. Weinmann for critical
readings of this manuscript. The Microarray Centre, which would like to thank J. Woodgett and
Neil Winegardin for technical input, was supported by ORDCF and Genome Canada. This work
was supported in part by Public Health Service grants CA22484, CA09135, and CA45240.
4
K. Noma, C. D. Allis, and S. I. Grewal, Science 293, 1150 (2001).
5
M. D. Litt, M. Simpson, M. Gaszner, C. D. Allis, and G. Felsenfeld, Science 293, 2453 (2001).
6
C. H. Bassing, W. Swat, and F. W. Alt, Cell 109 (Suppl.), S45 (2002).
336 immunochemical assays of chromatin functions [22]
Chromatin Immunoprecipitation
The basic procedures used for ChIP have been developed by and repre-
sent a fusion of contributions made by a large number of researchers. A
handful of these pioneering studies are outlined here: In 1981, whole cell
formaldehyde fixation was shown to crosslink and preserve chromatin
structure7,8; a few years later, antisera specific for RNA polymerase9 and
topoisomerase I10 were used to precipitate UV crosslinked protein:DNA
complexes; in 1988, histone antibodies were used in ChIP experiments11
and a correlation between core histone acetylation and transcriptionally
poised chromatin was observed.12 Recently, with the development and sub-
sequent commercialization of highly specific antibodies recognizing various
histone modifications, ChIP has become a very popular method for analyz-
ing chromatin structure in vivo. Although many variations of the basic
ChIP method are currently used, the backbone of our protocol relies
largely on the method reported by Kuras and Struhl13 and is described in
detail.
Procedure
1. The number of cells harvested for a single ChIP can vary
tremendously depending on the total number of DNA sites being
analyzed. The DNA we recovered from a single ChIP experiment was
used to examine the association of various histone modifications at ap-
proximately 30 loci across a large genomic region. However, the total
number of cells used per ChIP can be reduced if the analysis involves only
a few DNA sites. A transformed recombinase-deficient Pro B cell line was
chosen as the most practical line for examining the chromatin structure of
the murine IgH locus. However, this ChIP method has also been
successfully applied to additional suspension cell lines as well as two
non-adherent fibroblast cell lines.
7
V. Jackson and R. Chalkley, Proc. Natl. Acad. Sci. USA 78, 6081 (1981).
8
V. Jackson and R. Chalkley, Cell 23, 121 (1981).
9
D. S. Gilmour and J. T. Lis, Proc. Natl. Acad. Sci. USA 81, 4275 (1984).
10
D. S. Gilmour, G. Pflugfelder, J. C. Wang, and J. T. Lis, Cell 44, 401 (1986).
11
M. J. Solomon, P. L. Larsen, and A. Varshavsky, Cell 53, 937 (1988).
12
T. R. Hebbes, A. W. Thorne, and C. Crane-Robinson, EMBO J. 7, 1395 (1988).
13
L. Kuras and K. Struhl, Nature 399, 609 (1999).
[22] chromatin immunoprecipitation 337
Use approximately 2 108 cells for a single IP. To assess the efficiency
of the ChIP experiment, a Control IP lacking an antibody is performed in
parallel with one containing the specific antibody of interest. Therefore, for
a properly controlled experiment, approximately 4 108 cells are needed.
From a larger cell preparation, such as 8 108 cells, three independent
ChIP experiments can be performed with a single Control IP. Harvest
the cell population (collect adherent cells by trypsinization and wash the
cell pellet twice in ice-cold 1 PBS) and pellet at 640g for 4 min.
Resuspend the cell pellet thoroughly in 20 ml of culture media.
2. Fix the cells by adding formaldehyde to a final concentration of 1%
directly to the cell suspension. Allow the fixation reaction to proceed for
10 min in a fume hood at room temperature. Invert the suspension every
few minutes to mix the cells. Add 1/20 the volume of 2.5 M glycine to stop
the fixation reaction, and incubate for an additional 5 min at room
temperature. Transfer the cell suspension to an ice bucket.
3. Pellet the cells at 640g for 4 min (all centrifugations should be
performed at 4 until otherwise noted). Discard the supernatant into an
appropriate receptacle designated for formaldehyde waste. Resuspend the
cell pellet in 10 ml of 1 PBS and collect the cells by centrifugation.
Discard the supernatant and repeat the 1 PBS wash to remove all traces
of formaldehyde from the cell suspension
4. Isolate nuclei from fixed cells by washing the cell pellet three times
with 10 ml of cold lysis buffer (10 mM Tris-HCl, pH 7.5, 10 mM NaCl,
3 mM MgCl2, and 0.5% NP-40). After each wash, pellet the cells at 300g
for 5 min.
5. Resuspend the pelleted nuclei in 3 ml of MNase reaction buffer
(10 mM Tris-HCl, pH 7.5, 10 mM NaCl, 3 mM MgCl2, 1 mM CaCl2, 4%
NP-40, and 1 mM PMSF added just prior to use). Add 50 units of MNase
(Sigma) to the nuclei suspension, mix, and incubate for 10 min at 37 . The
amount of enzyme and the length of incubation should be specifically
tailored for optimum digestion when using different cell types or cell
numbers. Add EGTA to a final concentration of 3 mM to stop the
digestion reaction. At this point, save a 10-l aliquot and store at 20 .
This aliquot should be decrosslinked (described later) at the end of this
procedure and then run out on an agarose gel, providing a clear indication
of the efficiency of MNase digestion.
6. Add 60 l of 50 mM PMSF, 3 l of 2 mg/ml aprotinin, 3 l of
2 mg/ml leupeptin, 3 l of 2 mg/ml pepstatin, 300 l of 10% SDS, and
120 l of 5 M NaCl to the MNase-digested nuclei. Sonicate the nuclei to
lyse the nuclear membrane and further shear the chromatin into a sample
containing primarily 0.2–2 kb DNA fragments. The settings and conditions
used during sonication must be determined empirically and tend to vary
338 immunochemical assays of chromatin functions [22]
DNA Quantitation
A clear indication of the overall efficiency of the ChIP is assessed by
comparing the amount of DNA present in the IP sample to the amount
present in the Control IP. However, the amount of DNA recovered in both
of these samples is much too low for accurate quantitation using standard
procedures such as UV absorbance. The PicoGreen dsDNA Quantitation
Assay from Molecular Probes provides a rapid and sensitive technique
for quantitating DNA ranging from 25 pg/ml to 1 g/ml. Picogreen is a
cyanine dye that is essentially non-fluorescent when free in solution, but
exhibits a greater than 1000-fold enrichment in fluorescence upon binding
to DNA. The amount of DNA typically recovered in the Control IP is
approximately 20 ng. This amount should not vary drastically between in-
dependent ChIP experiments; however, the amount of DNA recovered
[22] chromatin immunoprecipitation 341
14
A. Giulietti, L. Overbergh, D. Valckx, B. Decallonne, R. Bouillon, and C. Mathieu,
Methods 25, 386 (2001).
15
D. G. Ginzinger, Exp. Hematol. 30, 503 (2002).
344 immunochemical assays of chromatin functions [22]
DNA, 10 pmol of each primer, and SYBR Green PCR Master Mix
(Applied Biosystems). Although a 50-l minimum reaction volume is
suggested for the Bio-Rad iCycler, we have consistently obtained
reproducible results with a 20-l reaction.
2. During amplification, the fluorescence signal intensity is plotted as a
function of the cycle number (see Fig. 2). After the reaction has finished, a
fluorescence signal cycle threshold (Ct) level must be set. The threshold
level should reside in the exponential phase of the real-time PCR reaction
and represents the point at which the Input sample can now be compared
to the IP sample. The number of cycles of each PCR reaction required to
reach the set threshold level is known as the Ct. Cycle threshold values are
directly proportional to the amount of DNA template present in each
reaction prior to amplification and provide the basis for calculating relative
fold enrichment values. The Ct values for each reaction within the
triplicate are averaged according to the following rule; if a Ct value within
the triplicate fluctuates greater than or less than 0.5 cycles in either
direction from the other two values in the triplicate (likely due to
pippetting errors), then this data point is not considered when calculating
the overall cycle threshold for that particular target sequence. Next, the
rate of amplification (R) should be calculated for each PCR reaction. Due
to the exponential nature of the PCR the theoretical optimum rate of
amplification is 2.0, but to make accurate and meaningful comparisons
between primer pairs, the differences in amplification rates should be
calculated. PCR primers that have been designed according to the rules
stated earlier typically have R values ranging between 1.85 and 2.0. These
values remain relatively constant for a particular primer pair, although on
rare occasions, variability in the rate is seen, which is likely due to
evaporation of a particular reaction from the 96-well PCR plate. The R
values for each reaction within a triplicate are averaged, except for those
individual R values that are less than 1.75 or greater than 2.25. If an R
value falls outside this bracket, then the data point is not used in the
calculation of overall primer amplification rate. The fold enrichment (see
mock calculation illustrated in Fig. 2) of a specific target sequence in the IP
sample relative to the Input sample is calculated using the following
equation:
Fold enrichment ¼ RðCtInput CtIP Þ :
346 immunochemical assays of chromatin functions [22]
Fig. 2. A graphic demonstration of the data generated by real-time PCR and the
calculations used in determining the enrichment of specific DNA sequences in an IP sample
relative to an Input sample. Fluorescence data generated in real time were plotted as a
function of the cycle number during PCR amplification of a particular target sequence from
representative IP and Input triplicate reactions. The cycle number is listed along the x-axis
and the fluorescence intensity is plotted on the y-axis. For this example, the Ct level was set to
a fluorescence of 80, which resides within the exponential phase of amplification. Extrapo-
lation of the fluorescence data from the set Ct level down to the x-axis provides Ct values for
both the IP and Input samples. Taking into account the rate of amplification (R), the fold
enrichment of a particular DNA sequence in the IP sample relative to the Input sample is
45.68.
16
J. R. Davie and V. A. Spencer, J. Cell Biochem. 32/33 (Suppl.), 141 (1999).
[22] chromatin immunoprecipitation 347
Fig. 3. The addition of MNase prior to sonication in the outlined ChIP method provides
the resolution needed to establish an enrichment pattern of the association of certain histone
modifications with genomic loci. (A) A schematic of the immunoglobulin heavy chain (IgH)
locus. Solid, black rectangles represent families of V segments, black vertical lines depict D
regions, and J gene segments are illustrated as hatched bars. The regulatory enhancer, E (E),
and the constant region, C3 (C), are depicted as a solid, black oval and a white rectangle,
respectively. Black dots located immediately above the schematic mark the location of each
PCR primer pair used in (B) and Fig. 4. (B) The enrichment of histone H3 (Lys9) methylation
across the IgH locus in Pro B cells as determined from multiple independent ChIP
experiments with and without the addition of MNase. Along the x-axis is the type of gene
segment amplified by each primer pair (V, D, J, E, and C). These segments are arranged from
left to right in the 50 to 30 orientation as they are located in the IgH locus. The fold enrichment
of each DNA sequence in an IP sample relative to an Input sample is shown.
Acknowledgments
We would like to acknowledge Joe Geisberg and other members of Kevin Struhl’s
laboratory for their help with the design of the ChIP protocol, PCR primers, and real-time
PCR analyses. We would also like to thank Laura Corey for her help with implementing
MNase into our protocol. We thank Mary Donohoe for her helpful discussions and critical
reading of the manuscript. This work was supported by NIH grant GM48026 to M.A.O.
[23] use of chromatin immunoprecipitation assays 349
Introduction
Systematic studies made possible by a variety of technological advances
are impacting many areas of biology. Of special relevance to the field of
chromatin are approaches that combine chromatin immunoprecipita-
tion (chromatin IP) with microarrays to analyze histone modifications
genome-wide in yeast.1,2 The resulting genomic maps provide a unique
global perspective on the functions of, and inter-relationships between,
different modifications.
Chromatin IP experiments use antibodies to immunoprecipitate a pro-
tein of interest and associated DNA from a solubilized chromatin prepar-
ation.3–7 Specialized antibodies can be used to enrich for DNA associated
with histones that exhibit specific post-translational modifications. A con-
siderable number of antibodies that recognize histones acetylated, methyl-
ated, or phosphorylated at specific residues have been developed, and
many are commercially available. Chromatin IP experiments typically
assess whether a particular gene, gene promoter, or genomic region is en-
riched in the IP sample relative to a whole cell extract (WCE) control. For
example, quantitative polymerase chain reaction (PCR) using pre-selected
primer pairs can be used to compare the representation of specific DNA
species in the IP and WCE. Although several regions can be queried simul-
taneously, these conventional studies are limited to a subset of genes or
regions chosen by the investigator.
Global studies have the potential to overcome selection bias by
analyzing comprehensively many or all elements in a system. Microarray
1
B. E. Bernstein, E. L. Humphrey, R. L. Erlich, R. Schneider, P. Bouman, J. S. Liu,
T. Kouzarides, and S. L. Schreiber, Proc. Natl. Acad. Sci. USA 99, 8695 (2002).
2
D. Robyr, Y. Suka, I. Xenarios, S. K. Kurdistani, A. Wang, N. Suka, and M. Grunstein, Cell
109, 437 (2002).
3
M. J. Solomon, P. L. Larsen, and A. Varshavsky, Cell 53, 937 (1988).
4
M. Braunstein, A. B. Rose, S. G. Holmes, C. D. Allis, and J. R. Broach, Genes Dev. 7,
592 (1993).
5
V. Orlando, H. Strutt, and R. Paro, Methods 11, 205 (1997).
6
M. H. Kuo and C. D. Allis, Methods 19, 425 (1999).
7
A. Hecht and M. Grunstein, Methods Enzymol. 304, 399 (1999).
technology, typically used for mRNA analysis,8,9 can be adapted for the
comprehensive analysis of chromatin IP samples.10,11 These studies rely
on several innovations, including:
1. The development of molecular biology protocols to amplify and label
chromatin IP DNA samples in a sequence independent manner and
2. The production of DNA microarrays that contain the open reading
frames (ORFs) as well as the regulatory regions
Recently this approach has been used to map relative levels of histone
acetylation and methylation genome-wide in yeast, and to identify regions
deacetylated by the various histone deacetylase enzymes in yeast (see
Fig. 1).1,2
Protocols
The protocol for analyzing histone modifications genome-wide in yeast
is presented in four sections:
1. Chromatin IP to isolate DNA associated in vivo with modified
histones
2. DNA amplification and labeling
3. Microarray hybridization and data acquisition
4. Data analysis and interpretation
Solutions
YPD: Yeast extract/peptone/dextrose
PBS: Phosphate-buffered saline pH 7.4
Buffer L: 50 mM HEPES-KOH, pH 7.5, 140 mM NaCl, 1 mM
EDTA, 1% Triton X-100, 0.1% sodium deoxycholate.
For 5 ml of buffer L, 20 l of protease inhibitor cocktail
(Sigma), and one mini complete protease inhibitor tablet
(Boeringer Mannheim) are added just before use
8
J. L. DeRisi, V. R. Iyer, and P. O. Brown, Science 278, 680 (1997).
9
L. Wodicka, H. Dong, M. Mittmann, M. H. Ho, and D. J. Lockhart, Nat. Biotechnol. 15,
1359 (1997).
10
V. R. Iyer, C. E. Horak, C. S. Scafe, D. Botstein, M. Snyder, and P. O. Brown, Nature 409,
533 (2001).
11
B. Ren, F. Robert, J. J. Wyrick, O. Aparicio, E. G. Jennings, I. Simon, J. Zeitlinger,
J. Schreiber, N. Hannett, E. Kanin et al., Science 290, 2306 (2000).
[23] use of chromatin immunoprecipitation assays 351
A 180-ml culture of yeast (Saccharomyces cerevisiae) is grown in YPD at 30
to an OD600 of 1.0. To cross-link proteins and DNA, 4.9 ml of 37% formal-
dehyde is added and the culture incubated at room temperature for 15 min
with occasional shaking. The formaldehyde is quenched by the addition of
9 ml of 2.5 M glycine, and the cells are left at room temperature for an ad-
ditional 5 min. The yeast are pelleted by centrifugation at 2000g for 5 min
at 4 , and washed twice with 180 ml ice-cold PBS. Washed cells are divided
into four aliquots, pelleted, frozen in liquid nitrogen, and stored at 80 .
To lyse cells, one aliquot (representing 45 ml of the original culture) is
resuspended in 400 l of buffer L. Five hundred microliters of acid-washed
glass beads are added and the mixture is vortexed on high for 45 min at 4 .
With the lid of the microfuge tube open, the bottom of the tube is punc-
tured with a hot 25-gauge needle and the cell lysate is spun to a new tube
(5 s full speed, cut off the top of the punctured tube before spinning). The
total volume of lysate is then adjusted to 700 l by addition of buffer
L. Chromatin is fragmented by sonication using a Branson 250 Sonifier
(4 20 s at setting 3 and 70% duty, with 20-s rests), and the insoluble
fraction precipitated by centrifugation at 14,000g for 15 min at 4 . The
supernatant contains the solubilized chromatin and is referred to as the
WCE. The size distribution of the chromatin fragments can be examined
by gel electrophoresis (see Fig. 2). A desired distribution is achieved by
varying sonication time.7
Prior to immunoprecipitation, 15 l of the WCE are set aside for a con-
trol. The remaining WCE is added to 50 l of packed protein A beads (pre-
equilibrated in buffer L) and rotated at 4 for 1 h to remove proteins that
bind nonspecifically to the beads. The supernatant is then transferred to an-
other tube that contains 20–30 l of anti-modified histone antibody and ro-
tated at 4 for 4 h. Antibody-histone complexes are precipitated by the
addition of 50 l of pre-equilibrated protein A beads and rotation at 4
for 1 h. The beads are collected by centrifugation at 3000g for 30 s and
washed successively with 1 ml of buffers L (twice), W1 (twice), W2 (twice),
and TE (twice) for 5 min each. The beads are then incubated in 125 l of
elution buffer at 65 for 10 min with frequent mixing. The beads are
pelleted by centrifugation at 10,000g for 2 min and the supernatant is
retained. After a second elution, the eluted samples are combined, overlaid
with mineral oil, and left at 65 overnight to reverse cross-links. In parallel,
15 l of WCE is combined with 235 l of elution buffer, overlaid, and
incubated at 65 overnight.
To isolate DNA, the IP sample and WCE control are incubated with an
equal volume of proteinase K solution at 37 for 2 h, extracted with 1
volume phenol (twice), 1 volume chloroform/isoamyl alcohol (25:1) (once),
and ethanol precipitated. The samples are centrifuged at 15,000g for
[23] use of chromatin immunoprecipitation assays 353
Fig. 2. WCE starting material and products amplified by R-PCR or IVT. Lanes 2–4 of a
2% non-denaturing agarose gel contain 200 ng nucleic acid. Lanes 1 and 5: 500 ng 100 bp
ladder (NEB). Lane 2: Yeast WCE. Lane 3: DNA amplified by R-PCR from WCE. Lane 4:
RNA amplified by IVT from WCE.
15 min at 4 , washed with cold 70% ethanol, dried, resuspended in 20 l of
TE with 10 g of RNase A, and incubated at 37 for 1 h. Prior to amplifi-
cation, the DNA is further purified with a MinElute PCR Purification
Kit (Qiagen).
12
S. K. Bohlander, R. Espinosa III, M. M. Le Beau, J. D. Rowley, and M. O. Diaz, Genomics
13, 1322 (1992).
13
J. D. Lieb, X. Liu, D. Botstein, and P. O. Brown, Nat. Genet. 28, 327 (2001).
354 immunochemical assays of chromatin functions [23]
the genomic DNA is copied twice using a primer adaptor that has a
conserved 50 sequence and a degenerate 30 sequence that anneals to the
ends of the genomic DNA. The copied DNA is amplified by PCR using
the 50 conserved sequence for priming. Both methods have been used to
amplify DNA obtained by chromatin IP.
Due to its exponential kinetics, PCR is particularly susceptible to
sequence- and length-dependent biases. Recently, a linear amplification
strategy has been applied to the amplification of genomic DNA.14 In this
procedure, a T7 promoter sequence is added to the ends of the genomic
DNA, creating a template suitable for in vitro transcription (IVT). Al-
though this IVT protocol has the added complication of working with
RNA, it preserves the species representation of the chromatin IP DNA
more effectively than the PCR methods. Here we present the R-PCR
and IVT amplification protocols as we have applied them in studies of
histone modifications.
DNA from the chromatin IP sample and the WCE control are treated
with calf intestinal alkaline phosphatase (CIP) to remove 30 phosphate
groups prior to IVT. DNA (up to 50 ng), 1 l NEB Buffer 3, and 0.25 l
of 10 U/l CIP (NEB) are combined and H2O is added to a total
volume of 10 l. The reaction is incubated at 37 for 1 h, and cleaned up
with a MinElute Reaction Cleanup Kit (Qiagen). The kit is used after sev-
eral steps in this protocol, always according to manufacturer’s instructions
except with the elution volume increased to 20 l. Terminal transferase
(TdT) is used to add 30 tails of relatively uniform length to the ends of
the genomic DNA. Ten microliters of CIP-treated DNA, 4 l TdT Buffer,
3 l 5 mM CoCl2, 1 l 8% nucleotide solution, and 2 l TdT enzyme
(10 U/l) are combined, overlaid with mineral oil, and incubated at 37
for 20 min. The reaction is halted by the addition of 2 l 0.5 M EDTA,
pH 8.0, and subjected to the MinElute Reaction Cleanup.
T7 promoters are incorporated at the ends of the DNA fragments as
follows: 20 l tailing reaction product, 0.6 l T7-A18B primer, 5 l 10
EcoPol buffer, 2 l 5.0 mM dNTPs, and 20.4 l nuclease-free water are
combined. In a thermal cycler, the samples are incubated at 94 for 2 min,
ramped down at 1 per second to 35 , held at 35 for 2 min, ramped down
at 0.5 per second to 25 and held while 2 l 5 U/l Klenow (NEB) is
added. The reaction is then incubated at 37 for 90 min, halted by the ad-
dition of 5 l 0.5 M EDTA, pH 8, and subjected to a MinElute Reaction
Cleanup. (Note: mineral oil is not used at this stage.)
Prior to the IVT, the volume is reduced to 8 l in a vacuum centrifuge
at medium heat. IVT is carried out using the T7 Megascript Kit (Ambion).
The samples are purified using an RNeasy Mini Kit (Qiagen) per manufac-
turer’s RNA cleanup protocol, except with an additional 500-l wash with
buffer RPE. RNA is quantified by absorbance at 260 nm and visualized on
a 2% native TAE agarose gel. Yields range from 10 to 60 g. The size
distribution should approximate that of the WCE (see Fig. 2).
Equal quantities of RNA from the IP sample and WCE control (2–4
g) are reverse-transcribed to generate DNA probes. In a 14.5-l
volume, the RNA is primed by addition of 1 l priming mix, incubation
at 70 for 10 min, and incubation on ice for 10 min. Then, 6 l 5 RT
buffer, 3 l 0.1 M DTT, 3 l nuclease-free H2O, 0.8 l 50 aa-dNTP
mix, and 2.5 l 200 U/l Superscript II (Invitrogen) are added, and the
mix is incubated at 42 for 2 h. The reaction is halted by the addition of
10 l 0.5 M EDTA, pH 8.0, and the RNA is hydrolyzed by the addition
of 10 l 1 N NaOH and incubated at 65 for 15 min. The solution is neu-
tralized with 25 l 1 M HEPES, pH 7.5. The entire samples are used for
labeling and hybridization.
[23] use of chromatin immunoprecipitation assays 357
Probe Labeling
Amplified samples are fluorescently labeled using monofunctional
Cy5 and Cy3 fluorescent dyes as described by Carroll et al.,15 and at:
www.microarrays.org. These derivatized dyes covalently couple to the ami-
noallyl groups incorporated in Round B of R-PCR, or in the RT step
following IVT. DNA samples are diluted with 400 l of nuclease-free
water, concentrated to 70 l in a YM-30 microcon (Millipore) by centri-
fugation at 11,000g, re-diluted and re-concentrated. After a final dilution,
the samples are concentrated to a volume of 18–20 l (microcons are pre-
weighed, and final volume determined by weight). Samples are collected by
inverting the microcon into a new microfuge tube and spinning at 4000g for
5 min. One microliter of 1 M NaHCO3, pH 9, is added and the samples are
transferred to a microfuge tube containing a dried aliquot of NHS cyanine
dye (Cy5 for IP sample; Cy3 for WCE control) and incubated at room tem-
perature for 75 min in the dark. [The dye aliquots are prepared by dissolv-
ing one vial of monofunctional Cy5 or Cy3 dye (Amersham) in 20 l
DMSO, distributing the solution into 10 microfuge tubes, and drying in a
vacuum centrifuge under low heat.] Reactions are diluted with 70 l H2O
and purified using a QIAquick PCR Purification Kit.
15
A. S. Carroll, A. C. Bishop, J. L. DeRisi, K. M. Shokat, and E. K. O’Shea, Proc. Natl. Acad.
Sci. USA 98, 12578 (2001).
16
A. P. Gasch, Methods Enzymol. 350, 393 (2002).
17
C. E. Horak and M. Snyder, Methods Enzymol. 350, 469 (2002).
358 immunochemical assays of chromatin functions [23]
Hybridization
The competitive hybridization used in DNA microarray analyses allows
flexibility in determining which samples to compare. In this presentation of
the protocol, the relative enrichment of a particular modification is ana-
lyzed genome-wide by comparing a chromatin IP sample to its WCE con-
trol. However, it is also possible to query changes in modification patterns
resulting from gene deletion or small-molecule treatment by comparing
two chromatin IP samples obtained from different sources using the same
antibody.
Reagents and solutions:
WS1: 387 ml Milli-Q water, 12 ml 20 SSC, 1 ml 10% SDS
WS2: 1 ml of 20 SSC in 399 ml of Milli-Q water
The following is adapted from the hybridization procedure described
at www.microarrays.org. Cy5-labeled chromatin IP probe and Cy3-labeled
WCE probe are combined, and concentrated in a microcon YM-30 filter
until 50 l of solution remain on the filter. The sample is collected by
inverting the microcon into a new microfuge tube and spinning at 4000g
for 5 min. Six microliters of 20 SSC, 3 l of 10 g/l poly(A), and
0.96 l of 1 M HEPES, pH 7.0, are added and the solution is filtered with
a pre-wet Millipore 0.45- filter spun at 12,000 rpm for 2 min. Ten percent
SDS, 0.9 l, is added, the mixed probe is placed in a 100 heat block for
2 min, placed at room temperature for 10 min, and then applied to the mi-
croarray. Slides printed with ORFs and INTs are placed in Telechem hy-
bridization chambers with the arrays facing up. One clean lifterslip (Erie
Scientific) is placed on each array so that the white strips face down. Half
of the mixed probe (about 20 l) is applied to the ORF array and the other
half to the INT array. Each well of the hybridization chambers is filled
with 40 l of 3 SSC, the chambers are then sealed and placed in a 60
water bath for 12–15 h.
Four Wheaton glass tanks containing WS1 or WS2 are prepared. A
metal slide rack is placed in the first tank of each solution. The microarray
slides are removed from the hybridization chambers and turned upside
down in the first tank of WS1. Slides are tilted to allow the lifterslips to fall
off, placed in the slide rack, and plunged up and down 20 times. This
dunking is repeated in the second tank of WS1 and in the two WS2 tanks.
The slides are dried by centrifugation at 1000 rpm in a Beckman tabletop
centrifuge for 2 min.
[23] use of chromatin immunoprecipitation assays 359
18
C. J. Roberts, B. Nelson, M. J. Marton, R. Stoughton, M. R. Meyer, H. A. Bennett, Y. D. He,
H. Dai, W. L. Walker, T. R. Hughes et al., Science 287, 873 (2000).
19
J. Gollub, C. A. Ball, G. Binkley, J. Demeter, D. B. Finkelstein, J. M. Hebert, T. Hernandez-
Boussard, H. Jin, M. Kaloper, J. C. Matese et al., Nucleic Acids Res. 31, 94 (2003).
360 immunochemical assays of chromatin functions [23]
20
S. Tavazoie, J. D. Hughes, M. J. Campbell, R. J. Cho, and G. M. Church, Nat. Genet. 22, 281
(1999).
21
M. B. Eisen, P. T. Spellman, P. O. Brown, and D. Botstein, Proc. Natl. Acad. Sci. USA 95,
14863 (1998).
22
B. E. Bernstein, J. K. Tong, and S. L. Schreiber, Proc. Natl. Acad. Sci. USA 97, 13708
(2000).
23
A. S. Weinmann, P. S. Yan, M. J. Oberley, T. H. Huang, and P. J. Farnham, Genes Dev. 16,
235 (2002).
24
B. Ren, H. Cam, Y. Takahashi, T. Volkert, J. Terragni, R. A. Young, and B. D. Dynlacht,
Genes Dev. 16, 245 (2002).
[24] sequential chromatin immunoprecipitation from animal tissues 361
1
D. F. Clayton, A. L. Harrelson, and J. E. Darnell, Jr., Mol. Cell Biol. 5, 2623 (1985).
2
M. Parrizas et al., Mol. Cell Biol. 21, 3234 (2001).
3
D. Chaya, T. Hayamizu, M. Bustin, and K. S. Zaret, J. Biol. Chem. 24 (2001).
experience in our laboratory and others has shown that they represent a
good starting point for working with other tissue types.
Note that our method for chromatin fragmentation involves a partial
sonication step followed by a complete restriction digestion step. The re-
striction digest, with an enzyme that recognizes a frequently occurring
four-base sequence, allows the target DNA to be definitively trimmed to
a known length and therefore limits the final PCR analysis solely to the
defined sequence. This eliminates a concern arising when only sonication
is used, where the immunoprecipitated protein is crosslinked to target
sequences that are statistically but not precisely sized by the extent
of DNA fragmentation, and hence may not be bound at the DNA site of
interest.
Another novel aspect of our approach is that we developed a method to
release antigen from the initial immunoprecipitation to allow a second, se-
quential immunoprecipitation of the same sample. This allows the investi-
gator to evaluate whether two distinct proteins were bound to the same
DNA fragment within the starting population of molecules. The resolution
of this approach is enhanced by the aforementioned restriction digestion of
chromatin prior to the first immunoprecipitation.
4
G. Pfeifer and A. D. Riggs, Genes Dev. 5, 1102 (1991).
[24] sequential chromatin immunoprecipitation from animal tissues 363
has further adapted it for pancreas, using pancreatic ducts, and it should be
adaptable to other organ systems containing vessels or ducts. In this pro-
cedure, the animal is sacrificed and the liver is perfused with phosphate-
buffered saline (PBS) and heparin to flush out blood cells. The chromatin
is then crosslinked in vivo by perfusing formaldehyde into the liver. In
order to avoid the body temperature of the animal to drop too quickly
during these steps and thereby inhibit the effect of formaldehyde, the per-
fusion is carried out on a heat block at 30 with 30 solutions. Liver nuclei
are then crudely purified.
The day before the experiment, prepare buffers A and B (see Table I)
but without 2-mercaptoethanol, which is added just prior to use. Store
buffers at 4 overnight, along with the following which are kept in the cold
room: pipettes, beakers, 15-ml Corex tubes, rotor and adaptors, slides
and cover slips, tissue homogenization device (dounce), cheesecloth, and
funnel. For tissue dissection, prepare: ethanol, 4 pins, big and small sharp
scissors, big and small tweezers, string, cotton-tipped applications, catheter
TABLE I
Solutions for Tissue Perfusion Protocol
Final concentrations
HEPES pH 7.6 1 15 mM 15 mM
KCl 3 60 mM 60 mM
NaCl 4 15 mM 15 mM
EDTA 0.5 0.2 mM 0.1 mM
EGTA 0.5 0.5 mM 0.1 mM
Sucrose 2.5 0.34 M 2.1 M
2-Mercaptoethanol 2 0.15 mM 0.15 mM
Glycine 2.5 125 mM —
Formaldehyde [20% in water, stored at 20 (Fischer Scientific, ref. F79–500)]
Sonication buffer
Stock Final concentration
Tris, pH 8.1 1M 50 mM
EDTA 0.5 M 2 mM
N-Lauroylsarcosine 10% 0.5% (filtered through 0.2 m)
PMSF 50 mM 0.5 mM
Protease inhibitors 5 mg/ml Each 5 g/ml (leupeptin,
trypsin inhibitor, antipain)
364 immunochemical assays of chromatin functions [24]
(Becton Dickinson, Angiocath, ref. 381137, 20 GA, 1.88 IN, 1.1 48 mm), two
10-ml syringes, plastic wrap, paper towels, small beaker, heat block,
aluminum-wrapped cardboard, hybridization oven at 30 , and a timer.
Prepare a beaker with 6 ml solution A, keep on ice. Place the following
solutions used for perfusion and pre-warm in the 30 oven: PBS with 1%
formaldehyde, PBS with 50 ng/ml heparin. Fill respective syringes with
10 ml of each solution. Prepare the dissecting area by placing heat block
in the center of the lab bench with the aluminum-wrapped cardboard cover
on top. Layer plastic wrap and then a paper towel above the cardboard.
Have all the tools ready on the side, including 10-cm long pieces of string.
Sacrifice one mouse at a time and place it on its back on the heat block.
Pin down the four legs and cut open the abdomen posterior to anterior. Cut
through the diaphram, damaging the least possible tissue to avoid bleeding.
Gently push the intestines on the side of the animal with a cotton-tipped
applicator in order to have better access to the liver. Seal the superior vena
cava, anterior to the liver, with a double knot using small tweezers and
string. Insert the catheter in the inferior vena cava, just above the kidney
junction. Remove the needle partway and blood should flow back into
the catheter. Cut the portal vein with small scissors. Screw the syringe con-
taining the PBS/heparin onto the needle and start very slowly perfusing the
liver (10 ml total). All the lobes of the organ should rapidly blanch white.
You should also see the buffer and blood flow out of the liver through the
portal vein. If the animal is sick or too old (over 3 months), the perfusion
will be difficult. If the lobes do not fully blanch, massage gently the liver
with a cotton-tipped applicator and keep perfusing very slowly. Switch
to the formaldehyde-containing syringe that is sitting in the 30 oven and
retrograde perfuse 10 ml slowly. The perfusions should take about
5–10 min. Let the formaldehyde sit in the perfused liver for 5 min in the
animal while on the heat block at 30 . Dissect out the liver and place in a
beaker in the 30 oven for another 5 min. Place the liver in 6 ml buffer A on
ice. The cold as well as the glycine will stop the crosslinking reaction. Mince
the liver with small scissors and let sit on ice while perfusing other livers.
Repeat this retrograde perfusion procedure on one or two more animals.
In the cold room, pour minced livers into a cold douncing apparatus on
ice. Rinse beaker with 5 ml buffer A and add to dounce. Aspirate excess
buffer from dounce. Dounce mechanically 12 times. Check a sample under
a phase-contrast microscope for quantitative release of nuclei and cells and
very few lumps of tissue. Filter through 8 layers of cheesecloth prewet with
buffer A. Layer gently onto a 1.5-ml cushion of a 1:1 mixture of buffers A
and B in 15-ml Corex tubes (if the tissue lysate volume is 4–6 ml, use one
tube; if 6–12 ml, use 2 tubes). Centrifuge in fixed-angle rotor (e.g., Sorvall
SS-34) at 10,000 rpm for 10 min at 4 . Gently decant the supernatant.
[24] sequential chromatin immunoprecipitation from animal tissues 365
TABLE II
Solutions for Isolating Nuclei
Final concentrations
HEPES, pH 7.6 1 15 mM 15 mM 15 mM
KCl 3 60 mM 60 mM 60 mM
NaCl 4 15 mM 15 mM 15 mM
EDTA 0.5 0.2 mM — —
EGTA 0.5 0.5 mM — —
Sucrose 2.5 0.34 M 2.1 M 0.34 M
2-Mercaptoethanol 2 0.15 mM 0.15 mM 0.15 mM
Spermine 0.5 0.15 mM — —
Spermidine 0.5 0.5 mM — —
MgCl2 1 — — 2 mM
Glycine 2.5 (kept at room temperature)
Formaldehyde 20% in water, stored at 20
366 immunochemical assays of chromatin functions [24]
minced livers into the cold dounce on ice. Rinse the beaker with 5 ml solu-
tion A and add to dounce, to get all tissue fragments. After the tissue
settles, aspirate excess supernatant from the dounce. Dounce 10 times with
a motor-driven pestle. Check under the microscope that you have nuclei
and cells and very few lumps of tissue. Dounce further accordingly. Filter
through 8 layers of cheesecloth prewet with solution A. Layer gently onto
a 1.5-ml cushion of a 1:1 mixture of buffers A and B in 15-ml Corex tubes
(if the tissue lysate volume is 4–6 ml, use one tube; if 6–12 ml, use 2 tubes).
Centrifuge in fixed-angle rotor (e.g., Sorvall SS-34) at 10,000 rpm for
20 min at 4 . Gently decant the supernatant. Resuspend the pellets gently
in 5–10 ml of solution C. Measure the volume. Add formaldehyde (20%) to
a final concentration of 1% and incubate for 10 min at 30 . Mix gently
every 2 or 3 min. Stop the crosslinking reaction by adding room tempera-
ture glycine to a final concentration of 125 mM and let incubate on ice for
5 min. Layer onto a 1.5-ml cushion of a 1:1 mixture of solutions A and B.
Centrifuge in fixed-angle rotor at 5000 rpm for 10 min at 4 . Gently decant
the supernatant, resuspend the pellet into 2–5 ml sonication buffer, and
follow the protocol for nuclear sonication.
Nuclear Sonication
These conditions are optimized for chromatin that has been crosslinked
within tissue and should be reduced for chromatin that has been cross-
linked within isolated nuclei. Nuclear pellets that were resuspended into
2–5 ml of sonication buffer should be incubated for 5 min at room tem-
perature, then 5 min on ice. Sonicate with a Branson Sonifier 250, mounted
with a microtip, using duty cycle constant, output #2. Pulse alternatively for
30 and 15 s, incubating 30 s on ice between pulses. Avoid bubbles and be
[24] sequential chromatin immunoprecipitation from animal tissues 367
careful not to heat the sample during the sonication, as the heat will reverse
the crosslinking in presence of Sarkosine.
TABLE III
Solutions for Preparing Antibody Covalently Bound to Protein A Beads,
Chromatin Immunoprecipitation, and DNA Purification
Tris, pH 8 1M 50 mM
NaCl 4M 150 mM (or 500 mM for I.P. washes)
IGEPAL CA-630 100% 1% (ICN)
Deoxycholic acid (Na salt) 10% 0.5% (Sigma)
SDS 20% 0.1%
RIPA buffer 2 without Tris
NaCl 4M 300 mM
IGEPAL CA-630 100% 2% (ICN)
Deoxycholic acid (Na salt) 10% 1% (Sigma)
SDS 20% 0.2%
TNESK 5X
Tris, pH 7.5 1M 50 mM
NaCl 4M 500 mM
EDTA 0.5 M 5 mM
SDS 20% 5%
Proteinase K 20 mg/ml 0.5 mg/ml (added just before use)
[24] sequential chromatin immunoprecipitation from animal tissues 371
Chromatin Immunoprecipitation
For each immunoprecipitation, use about 50–100 g of chromatin
DNA. Dialyze the total amount of purified crosslinked chromatin over-
night at 4 in 3500 cut-off dialysis bags against 500 ml RIPA buffer (with-
out IGEPAL). The next morning, dialyze a few hours against RIPA buffer
containing IGEPAL, then transfer to a polypropylene tube. To block non-
specific sites on the chromatin, add 2 l preimmune serum per milligram
chromatin and 200 l of 1:1 precleared protein A:RIPA. Incubate rotating
for 3 h at 4 . Centrifuge at 2500 rpm for 1 min at 4 and save supernatant
in fresh tube. Remove an aliquot of chromatin as an ‘‘input’’ sample.
Dispense 50–100 g aliquots of supernatant chromatin and add relevant
antibodies. Examples of experimental conditions include: no antibodies,
2 l nonimmune serum (stock 130 mg/ml), 2 l specific antiserum, 40 l
preimmune serum coupled to protein A, 40 l antiserum coupled to pro-
tein A. Incubate on rotator overnight at 4 . Add 40 l 1:1 precleared pro-
tein A:RIPA for samples that are not coupled to protein A and incubate
rotating for 2 h at 4 . Centrifuge all samples at 2500 rpm for 1 min at 4 .
Save supernatant for DNA analysis. Wash beads 4 times with 20 volumes
RIPA buffer containing 500 mM NaCl; for each wash, add buffer, let sit
on ice for 10 min, mixing occasionally, centrifuge, discard supernatant.
For final wash, transfer suspended beads to a fresh tube to eliminate chro-
matin bound nonspecifically to the plastic. For a single immunoprecipita-
tion experiment, beads are finally resuspended in 1 volume of TE. For
sequential immunoprecipitation, chromatin will be eluted from the beads,
as described in the following.
Acknowledgment
Work on this project was supported by a grant from the NIH (GM47903).
1
L. M. Silver and S. C. R. Elgin, Proc. Natl. Acad. Sci. USA 73, 423 (1976).
2
C. Rodriguez-Alfageme, G. T. Rudkin, and L. H. Cohen, Proc. Natl. Acad. Sci. USA 73,
2038 (1976).
3
M. Jamrich, A. L. Greenleaf, and E. K. F. Bautz, Proc. Natl. Acad. Sci. USA 74, 2079
(1977).
4
B. E. Schwartz, J. K. Werner, and J. T. Lis, Meth. Enzymol. 376, 393 (2004).
Acknowledgment
Work on this project was supported by a grant from the NIH (GM47903).
1
L. M. Silver and S. C. R. Elgin, Proc. Natl. Acad. Sci. USA 73, 423 (1976).
2
C. Rodriguez-Alfageme, G. T. Rudkin, and L. H. Cohen, Proc. Natl. Acad. Sci. USA 73,
2038 (1976).
3
M. Jamrich, A. L. Greenleaf, and E. K. F. Bautz, Proc. Natl. Acad. Sci. USA 74, 2079
(1977).
4
B. E. Schwartz, J. K. Werner, and J. T. Lis, Meth. Enzymol. 376, 393 (2004).
the third instar larvae (see also Schwartz et al.4 for more information on
endoreduplication).
Differential packaging of the DNA along each chromosome arm results
in a distinct banding pattern. Some of the constrictions represent sites of
underreplication, but in most cases, the bands and the interbands differ
only in their compaction of the DNA and not in their level of polyteny.5
In 1935, Bridges drew a detailed map of each chromosome arm; more re-
cently photographic representations have been made.6 A physical map
has been established through the technique of in situ hybridization. This
technique has allowed a specific band or region within a band to be cor-
related with the location of a specific gene or repetitive sequence at a
resolution of ca. 30 kb.7
Immunological methods for studying the association of proteins with
polytene chromosomes have been used to address a variety of biological
questions. In wild-type flies, immunostaining has been used to determine
the global distribution of one or more proteins; colocalization studies have
been done to determine if a protein of interest might be in close association
or part of a multiprotein complex with other proteins.8 Immunological
staining has also been performed on a variety of Drosophila species to
monitor conservation of a chromosomal protein and detect changes in
its distribution through evolution. Much has also been gained by taking
advantage of various genetic tools. Fly lines with chromosome rearrange-
ments have been used to show when the localization of a protein reflects
local as opposed to global structural features, for example, proximity to
the chromocenter.9 Mutations in chromosomal proteins can be assessed
by immunofluorescent staining to monitor both the extent of DNA binding
and distribution of the mutant protein, and the impact on chromosome
organization and distribution of other proteins.10 Finally, Drosophila
melanogaster can be transformed using a transposable element (usually a
P-element) carrying a cloned DNA fragment, allowing experiments to
examine the protein complexes that associate with a given construct.
5
A. Sperier and P. Sperier, Nature (London) 307, 176 (1984).
6
G. Lefevre, in ‘‘The Genetics and Biology of Drosophila’’ (M. Ashburner and E. Novitski,
eds.), 1a, p. 31. Academic Press, London, 1976.
7
J. Gall and M. L. Pardue, Proc. Natl. Acad. Sci. USA 63, 378 (1969).
8
C. D. Shaffer, G. E. Stephens, B. A. Thompson, L. Funches, J. A. Bernat, C. A. Craig, and
S. C. R. Elgin, Proc. Natl. Acad. Sci. USA 99, 14332 (2002).
9
T. C. James, J. C. Eissenberg, C. Craig, V. Dietrich, A. Hobson, and S. C. R. Elgin, Eur. J.
Cell Biol. 50, 170 (1989).
10
G. Schotta, A. Ebert, V. Krauss, A. Fischer, J. Hoffman, S. Rea, T. Jenuwein, R. Dorn, and
G. Reuter, EMBO J. 21, 1121 (2002).
374 immunochemical assays of chromatin functions [25]
11
G. Cavalli and R. Paro, Science 286, 955 (1999).
12
L. M. Silver, C. E. C. Wu, and S. C. R. Elgin, in ‘‘Methods in Chromosomal Protein Research’’
(G. Stein, J. Stein and L. Kleinsmith, eds.), p. 151. Academic Press, New York, 1977.
13
L. H. Cohen and B. V. Gotchel, J. Biol. Chem. 246, 1841 (1971).
[25]
immunofluorescent staining of polytene chromosomes
375
Fig. 1. Diagram of isolating, squashing, and staining salivary gland polytene chromosomes. Reprinted from: Methods
Cell Biol. 35, 203 (1991), with permission from Academic Press.
376 immunochemical assays of chromatin functions [25]
the glands are damaged, the morphology of the chromosomes will suffer. If
the adhering fat body cannot be removed at this step without damaging the
glands, it can be removed later during acetic acid fixation. The paired saliv-
ary glands will either appear slightly milky or clear and are somewhat
wishbone-shaped. Incubate the excised glands in Cohen buffer for 8–
10 min. Incubation of glands in this solution with detergent allows for the
dissolution of cytoplasmic membrane structures. Poor morphology of the
chromosomes will result if an 8- to 10-min incubation is exceeded. Next,
the glands are incubated for 3–25 min in a formaldehyde fixative. Shorter
incubations can be used but may result in incomplete fixation and hence
loss of some chromosomal proteins, while longer incubations result in
chromosomes that are difficult to spread out well. The glands are then
transferred to 45% acetic acid and incubated for 3–60 min. During this in-
cubation, any adhering fat body may be removed. This acetic acid fixation
is necessary to attain good phase morphology of the spread polytene
chromosomes. The phase morphology will suffer if the glands are left
in the acetic acid for more than 1 h. This formaldehyde fixation
technique minimizes the extraction of chromosomal proteins while
maintaining good antigenicity.1
Reagents. All stock solutions should be kept at 4 unless otherwise
indicated (see Table I).
TABLE I
Cohen buffer
10 mM MgCl2 50 l 1M
25 mM sodium glycerol 3-phosphate 125 l 1 M, pH 7 ( 20 )
3 mM CaCl2 150 l 0.1 M
10 mM KH2PO4 250 l 0.2 M
0.5% NP40 250 l 10%
30 mM KCl 750 l 0.2 M
160 mM sucrose 1.00 ml 0.8 M ( 20 )
H2O 2.425 ml
(Can be kept at 4 for 2–3 days.)
Formaldehyde fixative
0.1 M NaCl 100 l 5M
2 mM KCl 50 l 0.2 M
10 mM NaH2PO4 50 l 1 M, pH 7
2% NP40 1.00 ml 10%
2% formaldehyde 270 l 37%
H2O 3.53 ml
(Must be made fresh daily. Use 37% formaldehyde stock within 6 months of purchase.)
[25] immunofluorescent staining of polytene chromosomes 377
TABLE II
the edges of the slide between the thumb and forefinger. Once the salivary
gland cells and nuclei break open, the arms of the chromosomes will spread
out. Turn the slide over so that the coverslip is now on top of the slide and
monitor the spreading of the chromosome arms by looking at them under a
phase-contrast microscope at 400. Mark the location of the squash with a
marker so that it may be easily found when looking under the fluorescent
microscope. If the chromosome arms have not spread sufficiently, place the
slide on the bench, coverslip up, and tap the coverslip with the eraser end of
a pencil. When the spreading of the chromosome arms is satisfactory, flat-
ten the chromosomes by firmly applying thumb pressure to the coverslip.
Place a folded tissue between the thumb and the coverslip so that grease
from the hands does not get on the slide. Be sure that the coverslip does
not move under pressure. If movement occurs, stretching and breaking of
the chromosome arms will result. If too much pressure is applied, the result
will be fragmented chromosomes. Also, be sure that the squash does not
dry out. If the squash is allowed to dry, subsequent staining will be poor.
Once the spread is satisfactory, quickly submerge the slide in liquid nitro-
gen using hemostatic forceps. Once bubbling has ceased, remove the slide
from the liquid nitrogen and flick off the coverslip with a razor blade. Be
sure not to scrape the slide or the chromosomes may be scraped off. Before
the specimen is allowed to thaw, immerse the slide in Tris-buffered saline
(TBS-Tween, see Table III). If the slides will not be used within 2–4 h, you
may store them in slide storage medium at 20 for up to a month. If the
squash is allowed to dry out at any time during the procedure, the result
will be poor staining of the chromosomes; the stain will outline the chromo-
somes rather than staining specific bands (see Fig. 2). The squashes are
TABLE III
10 TBS-Tween
0.2 M Tris-HCl 200 ml 1 M, pH 8
17% NaCl 170 g
Tween-20 (Sigma) 10 ml 100%
H2O Bring to 1 L
(Dilute 10-fold in H2O before use.)
Slide storage medium
67% glycerol 335 ml 100%
a
33% PBS 165 ml
a
One liter of phosphate-buffered saline (PBS) is made with 20 g of NaCl, 0.5 g of KCl,
0.5 g of KH2PO4, and 1.45 g of Na2HPO4.
[25] immunofluorescent staining of polytene chromosomes 379
most vulnerable to drying out after they are flattened. After flattening, the
liquid layer is very thin. If the acetic acid is seen to recede from the edges of
the coverslip, particularly at the corners, the specimen may be too dry to be
stained well. Therefore, during the squashing procedure, perform all steps
as quickly as possible without sacrificing the quality of your specimen. A
humidifier on the bench may help if the air is dry. Results will improve with
practice. A troubleshooting guide is provided at the end of this section to
help solve any problems that are encountered.
16
A. H. Coons, H. J. Creech, and R. N. Jones, Proc. Soc. Exp. Biol. Med. 47, 200 (1941).
17
A. H. Coons, H. J. Creech, R. N. Jones, and E. Berliner, J. Immunol. 45, 159 (1942).
18
M. Goldman, in ‘‘Fluorescent Antibody Methods.’’ Academic Press, New York, 1968.
19
L. A. Sternberger, in ‘‘Immunocytochemistry.’’ Prentice-Hall, Englewood Cliffs, NJ, 1974.
20
C. A. Williams and M. W. Chase, in ‘‘Methods in Immunology and Immunocytochemistry.’’
5. Academic Press, New York, 1976.
21
T. H. Weller and A. H. Coons, Proc. Soc. Exp. Biol. Med. 86, 789 (1954).
22
L. M. Silver and S. C. R. Elgin, in ‘‘The Cell Nucleus’’ (H. Busch, ed.), 5, p. 215. Academic
Press, New York, 1978.
[25] immunofluorescent staining of polytene chromosomes 381
TABLE IV
Mounting solution
90% glycerol 90 ml 100%
0.1 M Tris-HCl 10 ml 1 M, pH 7
0.2% n-propyl gallate 0.2 g
382 immunochemical assays of chromatin functions [25]
viewing and photography. It is helpful to scan the slide with a 10 dry lens
first to find a good squash and then move to a 400 oil immersion lens
for photographs. The images are processed in Photoshop and printed on
photo-quality glossy paper using a high-quality ink jet printer or laser
printer. Different fluorophores fade at different rates. It is advantageous
to minimize the amount of time spent viewing the chromosomes under
the fluorescent microscope prior to photography to minimize fading. It is
also advantageous to store the slides in the dark at 4 ; this keeps the slides
from drying out and the fluorophores from fading.
The methodology described thus far is for labeling one protein of inter-
est. Multiple proteins may be viewed on polytene chromosomes by incubat-
ing with the primary antibodies at the same time (mixing them together in
the diluent) or incubating sequentially with wash steps in between. The
secondary antibodies may also be added together or sequentially.
Pulverize n-propyl gallate with a mortar and pestle before weighing,
and allow it to dissolve by stirring overnight in the solution. n-Propyl
gallate inhibits the loss of fluorescence during viewing.
Troubleshooting Guide
The following is a troubleshooting guide. For each problem stated,
there is a possible explanation and an approach described to correct the
problem. (Adapted from Methods Cell Biol. 35, 214–216 (1991), with
permission from Elsevier.)
l. Chromosomes Are Present, but Not Coming into Focus or Are Blurry
1. Two coverslips may be stuck together on the slide.
Correction: Remove the upper coverslip with a razor blade.
2. Chromosomes were not squashed flat and are difficult to focus on
due to multiple focal planes.
Correction: Be sure to sufficiently flatten the chromosomes when
squashing.
3. The table that the microscope is on may be vibrating while you are
viewing the chromosomes and taking pictures.
Correction: A vibration isolation table may be needed if a suitable
stable surface cannot be found.
23
T. C. James and S. C. R. Elgin, Mol. Cell. Biol. 6, 3862 (1986).
24
T. C. James, J. C. Eissenberg, C. Craig, V. Dietrich, A. Hobson, and S. C. R. Elgin, Eur. J.
Cell Biol. 50, 170 (1989).
25
J. C. Eissenberg, T. C. James, D. M. Foster-Hartnett, T. Hartnett, V. Ngan, and S. C. R.
Elgin, Proc. Natl. Acad. Sci. USA 87, 9923 (1990).
26
D. A. R. Sinclair, R. C. Mottus, and T. A. Grigliatti, Mol. Gen. Genet. 191, 326 (1983).
27
J. B. Spofford, in ‘‘The Genetics and Biology of Drosophila’’ (M. Ashburner and
E. Novitski, eds.), 2a, p. 955. Academic Press, Orlando, FL, 1976.
28
A. J. Bannister, P. Zegerman, J. F. Partridge, E. A. Miska, J. O. Thomas, R. C. Allshire, and
T. Kouzarides, Nature 410, 120 (2001).
29
M. Lachner, D. O’Carroll, S. Rea, K. Mechtler, and T. Jenuwein, Nature 410, 116 (2001).
[25] immunofluorescent staining of polytene chromosomes 387
We are interested in other proteins that bind to HP1 and might contrib-
ute to a multiprotein complex required for heterochromatin-induced gene
silencing. Through a yeast two-hybrid screen using HP1 as bait, an HP1-
interacting protein, Heterochromatin Protein 2 (HP2), has been identified.8
Upon staining of polytene chromosomes with a polyclonal antibody pre-
pared against a peptide of HP2 from the C-terminal region, a polytene
chromosome staining pattern nearly coincident with that of HP1 is seen
(see Fig. 3). Mutations in HP2 result in suppression of PEV, suggesting that
HP2 also has a role in heterochromatin structure. The nearly coincident
staining pattern of these two proteins shows that they colocalize, thus
suggesting that they may be part of a multiprotein complex.
Methylation of lysine 9 on histone H3 has been found to recruit HP1.28,29
HP1 also interacts with the histone methyltransferase SU(VAR)3–9,10,30,31
providing a mechanism for the spreading of heterochromatin. Various
30
L. Aagaard, G. Laible, P. Selenko, M. Schmid, R. Dorn, G. Schotta, S. Kuhfittig, A. Wolf,
A. Lebersorger, P. B. Singh, G. Reuter, and T. Jenuwein, EMBO J. 18, 1923 (1999).
31
S. Rea, F. Eisenhaber, D. O’Carroll, B. D. Strahl, Z. W. Sun, M. Schmid, S. Opravil, K.
Mechtler, C. P. Ponting, C. D. Allis, and T. Jenuwein, Nature 406, 593 (2000).
388 immunochemical assays of chromatin functions [25]
Fig. 4. Immunfluorescent staining for HP1 versus histone H3 acetylated at lysine 14 on the
fourth chromosome of D. melanogaster. Panel a: HP1. Panel b: Histone H3 acetylated at
lysine 14. Notice the nonoverlapping staining patterns of HP1 and this histone modification,
suggesting different functions.
32
J. Nakayama, J. C. Rice, B. D. Strahl, C. D. Allis, and S. I. S. Grewal, Science 292, 110 (2001).
[25] immunofluorescent staining of polytene chromosomes 389
33
G. Reuter and I. Wolff, Mol. Gen. Genet. 182, 516 (1981).
390 immunochemical assays of chromatin functions [25]
34
Y. Li, J. R. Danzer, P. Alvarez, A. S. Belmont, and L. L. Wallrath, Development 130,
1817 (2003).
[25] immunofluorescent staining of polytene chromosomes 391
Fig. 6. (continued )
392 immunochemical assays of chromatin functions [25]
Conclusions
In conclusion, we have presented a method for determining the in vivo
distribution of chromosomal proteins on Drosophila polytene chromo-
somes. Some applications of this technique have also been discussed. By
combining genetic, biochemical, and molecular biology techniques
with the cytological approach, a greater understanding of the molecular
35
G. Cavalli and R. Paro, Cell 93, 505 (1998).
36
K. Ekwall, T. Olsson, B. M. Turner, and R. C. Allshire, Cell 91, 1021 (1997).
mechanisms of gene regulation can be gained. Few other systems offer the
well-developed genetic tools in combination with the ability to perform
cytological studies as Drosophila does. For more information on the
formation of polytene chromosomes and various other applications, see
Schwartz et al.4 this volume.
Background
Drosophila melanogaster was first recognized as a valuable experimen-
tal organism 100 years ago. One of the particularly attractive features of
this model system for studies of genes and their regulation is the ‘‘giant’’
or polytene chromosomes that occur in the secretory glands of Drosophila
(as well as other dipteran flies). Historically, polytene chromosomes pro-
vided an important link between the genetic map and the physical location
of deletions, insertions, inversions, and translocations on the genetic map.
More recently, these chromosomes have provided an efficient means of
analyzing both the global distribution of particular proteins on chromo-
somes and the recruitment of particular proteins to specific chromosomal
loci that are undergoing changes in activity.
Polytene chromosomes form in cells that grow in size without dividing
during larval development. In these cells, homologous diploid pairs of
chromosomes are tightly paired and undergo successive rounds of amplifi-
cation. Each synapsed pair may replicate up to nine times forming around
1000 strands of DNA, which remain aligned and attached to each other.
This precise alignment allows differences in chromatin compaction to be
seen as a series of bands and interbands extending across the width of the
chromosome arms. Electron microscopy studies show that about 95% of
the DNA is concentrated in the bands, which are much more compacted
regions of chromatin than the interbands, which separate the bands. Al-
though the packing ratio of chromatin in polytene chromosomes varies
regionally, the average packing ratio has been estimated to be between
30 and 57, about 100 times more extended than a metaphase chromosome.
This is slightly less compact than a 30-nm chromatin fiber, which has a
mechanisms of gene regulation can be gained. Few other systems offer the
well-developed genetic tools in combination with the ability to perform
cytological studies as Drosophila does. For more information on the
formation of polytene chromosomes and various other applications, see
Schwartz et al.4 this volume.
Background
Drosophila melanogaster was first recognized as a valuable experimen-
tal organism 100 years ago. One of the particularly attractive features of
this model system for studies of genes and their regulation is the ‘‘giant’’
or polytene chromosomes that occur in the secretory glands of Drosophila
(as well as other dipteran flies). Historically, polytene chromosomes pro-
vided an important link between the genetic map and the physical location
of deletions, insertions, inversions, and translocations on the genetic map.
More recently, these chromosomes have provided an efficient means of
analyzing both the global distribution of particular proteins on chromo-
somes and the recruitment of particular proteins to specific chromosomal
loci that are undergoing changes in activity.
Polytene chromosomes form in cells that grow in size without dividing
during larval development. In these cells, homologous diploid pairs of
chromosomes are tightly paired and undergo successive rounds of amplifi-
cation. Each synapsed pair may replicate up to nine times forming around
1000 strands of DNA, which remain aligned and attached to each other.
This precise alignment allows differences in chromatin compaction to be
seen as a series of bands and interbands extending across the width of the
chromosome arms. Electron microscopy studies show that about 95% of
the DNA is concentrated in the bands, which are much more compacted
regions of chromatin than the interbands, which separate the bands. Al-
though the packing ratio of chromatin in polytene chromosomes varies
regionally, the average packing ratio has been estimated to be between
30 and 57, about 100 times more extended than a metaphase chromosome.
This is slightly less compact than a 30-nm chromatin fiber, which has a
1
J. A. Simon, C. A. Sutton, R. B. Lobell, R. L. Glaser, and J. T. Lis, Cell 4, 805 (1985).
2
N. A. Winegarden, K. S. Wong, M. Sopta, and J. T. Westwood, J. Biol. Chem. 271, 2697 (1996).
3
L. M. Silver and S. C. Elgin, Chromosoma 68, 101 (1978).
4
U. Plagens, A. L. Greenleaf, and E. K. Bautz, Chromosoma 59, 157 (1976).
5
J. R. Bone and M. I. Kuroda, Genetics 144, 705 (1996).
6
T. C. James, J. C. Eissenberg, C. Craig, V. Dietrich, A. Hobson, and S. C. Elgin, Eur. J. Cell
Biol. 50, 170 (1989).
[26] immunofluorescent labeling of DROSOPHILA polytene chromosomes 395
7
E. D. Andrulis, J. Werner, A. Nazarian, H. Erdjument-Bromage, P. Tempst, and J. T. Lis,
Nature 420, 837 (2002).
8
S. Takada, J. T. Lis, S. Zhou, and R. Tjian, Cell 101, 459 (2000).
9
J. T. Lis, P. Mason, J. Peng, D. H. Price, and J. Werner, Genes Dev. 14, 792 (2000).
10
E. D. Andrulis, E. Guzman, P. Doring, J. Werner, and J. T. Lis, Genes Dev. 14, 2635 (2000).
11
B. E. Schwartz, unpublished results.
396 immunochemical assays of chromatin functions [26]
Fig. 1. Triple immuno-stain. Polytene chromosomes from heat shocked larvae were
stained with antibodies against HSF (red), the cyclin T subunit of P-TEFb (blue), and the
serine 2 phosphorylated form of RNA polymerase II (green).
Fig. 2. Elongating form of Pol II (green) resolves from promoter-bound HSF (red) during
heat shock. The heat shock puff from a single transgenic copy of an Hsp70-lacZ fusion gene
at chromosomal locus 9D is shown. The HSF band at the left edge of the puff marks
the promoter of the transgene, while elongating Pol II appears within the body of the puff.
new labeled band created at a transgenic site (not present in the parental
line) to which the gene or DNA segment has been moved provides
unambiguous proof of the assignment. Creating transgenic sites with
smaller and smaller segments can, in principle, pinpoint the DNA sequence
with which the protein associates.
The sensitivity of the immunofluorescent signal can be further en-
hanced by examining polymeric sequences introduced on transgenes. We
have used this approach to create a polymer of 40 copies of a 55-bp
hsp70 promoter fragment12 that vigorously recruits HSF during heat
shock.13 By co-immunostaining with antibodies against other factors, we
have identified proteins that interact with DNA-bound HSF (see Fig. 3).
Therefore, this technique not only allows the amplification of otherwise
weak immunofluorescent signals but also allows one to determine the
protein recruitment ability of defined promoter elements.
12
H. Xiao, Ph.D. Thesis, Cornell University, 1989.
13
L. S. Shopland and J. T. Lis, Chromosoma 105, 158 (1996).
398 immunochemical assays of chromatin functions [26]
Fig. 3. (A) Transgenes containing a polymer of an Hsp70 promoter fragment recruit HSF
(green) during heat shock. (B) HSF can, in turn, recruit TFIIH (red). (C) Merge.
[26] immunofluorescent labeling of DROSOPHILA polytene chromosomes 399
Methods
y
Fix Buffer: 50 l 37% paraformaldehyde (1.85 g paraformaldehyde in 5 ml H2O and 70 l 1
N KOH). Heat to 80 for 30 min. Store at 20 in aliquots:
450 l acetic acid
500 l H2O
Keeps for 2 h at room temperature
400 immunochemical assays of chromatin functions [26]
Staining of Slides
1. Remove slides from liquid nitrogen. Blow breath over coverslip.
Using a razor blade under one corner, flick the coverslip off
quickly. Using a diamond-tip pen, make small lines on the
microscope slide to indicate where coverslip was.
2. Wash:
2 10 min in PBS
1 10 min in PBS with 1% Triton.
3. Block 40–60 min in block solution (PBS with 5% non-fat milk
powder).
4. Rinse well in PBS.
5. Add primary antibodies in PBS and 1% BSA, for a total of 20 l
per slide. Put antibody solution directly over the area where the
squashes are. Cover with 22-mm2 coverslip, using lines etched in
the slide to position the coverslip. Leave overnight at 4 in humid
chamber.
{
Rain-X–treated coverslips: using a Kimwipe saturated with Rain-X, coat 22 mm2 coverslip
with Rain-X. Lay coverslip on Kimwipe to dry. Rinse coverslip with Kimwipe saturated with
water. Let it dry. Clean the coverslips with lens paper prior to using. These treated coverslips
can be stored and used later.
x
Base treated slides: dissolve 140 g NaOH in 560 ml dH2O. Add 840 ml 95% EtOH. Put
microscope slides in glass slide trays. Cover with NaOH solution. Leave shaking for 2 h. Do
4 1 h rinses with dH2O. Put in drying oven. The slides need to dry at least overnight before
using. The slides should be stored in the oven until used. Slides that are removed from the
oven and not used can be returned to the oven and used another time.
[26] immunofluorescent labeling of DROSOPHILA polytene chromosomes 401
14
G. Lefevre, Jr., in ‘‘The Genetics and Biology of Drosophila’’ (M. Ashburner and
E. Novitski, eds.), p. 31. Academic Press, New York, 1976.
15
M. Ashburner, in ‘‘Developmental Studies on Giant Chromosomes’’ (W. Beerman, ed.),
p. 101. Springer, Berlin, 1972.
16
V. Sorsa, ‘‘Chromosome Maps of Drosophila.’’ CRC Press, 1988.
[26] immunofluorescent labeling of DROSOPHILA polytene chromosomes 403
it is only possible to resolve the 50 from the 30 regions of a single gene with
certainty. This may be somewhat improved with a confocal microscope or
by taking a photographic Z-series and using software that can subtract
background fluorescence. If the goal is to map the position of a protein
within a gene with very high resolution, a more appropriate assay is
chromatin immunoprecipitation or ChIP, coupled with assays of the ChIP
DNA on microarrays. In this approach, proteins are crosslinked to chroma-
tin in living cells with formaldehyde, much as they are in polytene chromo-
some fixation. The protein of interest is immunoprecipitated. The DNA
covalently attached to the immunoprecipitated protein is purified, PCR
amplified, tagged with a fluorescent label, and used to probe a microarray
(chip) that contains DNA representing various portions of the gene of
interest. Microarrays containing spotted Drosophila cDNA libraries have
been produced (BD Biosciences, Palo Alto, CA); however, the complete
genome is not yet available.
Another limitation of polytene immunofluorescence analysis is the
challenge of quantifying immunofluorescent signals. The amount of immu-
nofluorescent signal at a particular locus can vary from squash to squash
for several reasons. First, the extent of polytenization can differ in chro-
mosomes derived from the same salivary gland. This could lead to exagger-
ated signals in over-replicated chromosomes and more modest signals in
comparatively under-replicated chromosomes or chromosomal regions.
Second, the preparation of the squash can lead to deformations such as
over-stretching and twisting of the chromosome arms—aberrations that
can in some cases significantly affect the measured amount of fluorescent
signal. Due to the nature of the squashing procedure, these deformations
almost always occur. It is therefore helpful to screen for damaged chromo-
somes with a general DNA stain such as DAPI or Hoechst (see also
Stephens et al.17 for more information on troubleshooting squashes).
Although quantifying immunofluorescent bands presents the above-
mentioned problems, it can be accomplished successfully if the chromo-
somes are carefully prepared and screened for those with crisp,
undistorted banding patterns. We have used NIH Image software to quan-
tify the intensity of signals at specified loci.13 By measuring the signals of
several squashes it is possible to make quantitative conclusions within
acceptable limits of error. Typically, these measurements are borne out
by other experimental approaches, such as ChIP analysis.
Not all chromosomal proteins show highly specific, strong-staining
patterns. While antibodies to many chromosomal proteins that have been
examined produced well-defined patterns of staining that can be readily
17
G. E. Stephens, C. A. Craig, Y. Li, L. L. Wallrath, and S. C. R. Elgin, Meth. Enzymol.
376, 372 (2004).
404 immunochemical assays of chromatin functions [26]
mapped, some do not (we estimate more than half of the proteins we tested
work well). There are those that show no staining pattern, even when mul-
tiple antibodies are used, and those that produce detectable weak staining
of specific sites and a diffuse staining of chromosomes. In some cases,
examining multiple high-quality chromosome squashes can produce slides
with sufficient signal to noise ratios to allow mapping.18 The range of re-
sponses seen with different proteins and different antibodies can be simply
explained in some cases by the quality of antibodies and abundance of a
protein on a particular site, but in other cases, other factors such as the
efficiency of fixing proteins to chromosomes, availability of the antigen to
chromosome surface, or the ability of antigenic determinants to survive fix-
ation will influence signal strength. Trying more than one slide preparation
protocol can sometimes help solve these problems.
The study of transcription in Drosophila is often aided by the use of
drugs that can inactivate or otherwise regulate the activity of specific
factors. Because the reliable uptake of such drugs by whole flies can pre-
sent dosage problems, it is worthwhile to note that explanted salivary
glands have been successfully treated with drugs in vitro. In this way, the
drug dosage can be precisely controlled and delivered directly to tissues
containing polytene chromosomes. Glands cultured in buffered media
remain physiologically active and are also capable of mounting a heat
shock response. Therefore, it is possible to inactivate or regulate a specific
transcription factor, apply a heat shock treatment or other inducement of
particular gene expression, prepare a chromosome spread, and immuno-
stain to determine the effect on transcription in vivo. This approach has
been used to show that heat shock gene transcription is sensitive to alpha
amanitin.19 The localization of other factors may be indirectly influenced
by inactivating the protein of interest, and this can also be assayed by
immuno-staining.
In summary, immunofluorescent staining of polytene chromosomes is a
reliable, fast, and convenient means of assessing the genome-wide distribu-
tion of chromatin binding proteins. The availability of transgenic flies and
the ability to treat salivary glands with drugs makes it a powerful tool to
investigate the function and dynamics of transcription and other factors
in vivo.
18
J. M. Park, J. Werner, J. M. Kim, J. T. Lis, and Y. J. Kim, Mol. Cell 8, 9 (2001).
19
J. L. Compton and B. J. McCarthy, Cell 14, 191 (1978).
[27] X-chromosome inactivation in mouse embryonic stem cells 405
1
M. F. Lyon, Nature, 190, 372 (1961).
2
P. Avner and E. Heard, Nat. Rev. Genet. 1, 59 (2001).
3
N. Brockdorff, Trends Genet. 18, 352 (2002).
4
N. Takagi, O. Sugawara, and M. Sasaki, Chromosoma 85, 275 (1982).
5
D. P. Norris, H. N. Brockdorff, and S. Rastan, Mamm. Genome, 1, 78 (1991).
6
P. Jeppesen and B. M. Turner, Cell 74, 281 (1993).
7
B. A. Boggs, P. Cheung, E. Heard, D. L. Spector, A. C. Chinault, and C. D. Allis, Nat.
Genet. 30, 73 (2002).
8
A. H. Peters, J. E. Mermoud, D. O’Carroll, M. Pagani, D. Schweizer, N. Brockdorff, and
T. Jenuwein, Nat. Genet. 30, 77 (2002).
9
J. Silva, W. Mak, I. Zvetkova, R. Appanah, T. B. Nestorova, Z. Webster, A. H. F. M. Peters,
T. Jenuwein, A. P. Otte, and N. Brockdorff, Dev. Cell 4, 481 (2003).
10
K. Plath, J. Fang, S. Mlynarczyk-Evans, R. Cao, K. A. Worringer, H. Wang, C. C. de la
Cruz, A. Otte, B. Panning, and Y. Zhang, Science 300, 131 (2003).
11
C. Costanzi and J. R. Pehrson, Nature 393, 599 (1998).
12
B. P. Chadwick and H. F. Willard, Hum. Mol. Genet. 10, 1101 (2001).
13
A. M. Keohane, L. P. O’Neill, N. D. Belyaev, J. S. Lavender, and B. M. Turner, Dev. Biol.
180, 618 (1996).
14
A. Wutz and R. A. Jaenisch, Mol. Cell 5, 695 (2000).
15
E. Heard, C. Rougeulle, D. Arnaud, P. Avner, C. D. Allis, and D. L. Spector, Cell 107, 727
(2001).
16
J. E. Mermoud, B. Popova, A. H. F. M. Peters, T. Jenuwein, and N. Brockdorff, Curr. Biol.
12, 247 (2002).
17
J. Chaumeil, I. Okamoto, M. Guggiari, and E. Heard, Cytogenet. Res. 75, (2002).
[27]
X-chromosome inactivation in mouse embryonic stem cells
Fig. 1. Examples of immunofluorescence combined with RNA FISH on interphase nuclei. Modifications of the N-terminal histone tails
of histone H3 on the inactive X chromosome in differentiating female ES cells. In each case, nuclei are shown with typical patterns observed
in undifferentiated female ES cells (panel A) or during differentiation (panel B). Immunodetection with Alexa GAR 568 conjugated
secondary antibody (red, column 2 of each panel) was combined with Xist RNA FISH (Spectrum Green labeled probe, green, column 3 of
each panel). Three antibodies were used here, which detect di-methylation of H3 lys-4 (from Upstate Biotechnology), di-methylation of H3
lys-9 (gift from D. Allis, also available from Upstate Biotechnology), and di-/tri-methylation of H3 lys-27 (gift from D. Reinberg, see Sarma
and Reinberg, this volume). Prior to the onset of X inactivation, in undifferentiated ES cells, Xist is transcribed at a low level from both X
chromosomes and the primary transcripts can be detected as a ‘‘dot’’ at each Xist locus (green, A). At the beginning of inactivation, Xist
407
RNA starts to accumulate, over the future inactive X chromosome, and can be detected as a green domain in the nuclei (green, B). In this
408 immunochemical assays of chromatin functions [27]
Cell Culture
18
A. Kuzmichev, K. Nishioka, H. Erdjument-Bromage, P. Tempst, and D. Reinberg, Genes
Dev. 16, 2893 (2002).
19
T. P. Rasmussen, M. A. Mastrangelo, A. Eden, J. R. Pehrson, and R. Jaenisch, J. Cell Biol.
150, 1189 (2000).
20
B. Hogan, R. Beddington, F. Constantini, and E. Lacy, ‘‘Manipulating the Mouse Embryo:
A Laboratory Manual.’’ Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY,
1994.
same window of time, some histone modifications begin to appear on the inactive X
chromosome (red, B). In merge images of immunofluorescence and RNA FISH (third
column, B), green coloration shows exclusion of the modification or yellow coloration shows
the enrichment on the Xist domain. The inactive X chromosome is depleted in H3 Lys-4
di-methylation (first row) and enriched in H3 Lys-9 di-methylation (second row) and Lys-27
di-/tri-methylation (third row) (B, arrowheads). DNA is stained with DAPI (blue, first column
of each panel).
[27] X-chromosome inactivation in mouse embryonic stem cells 409
cat. no. 10270106). It should be noted that only low passage number (no
higher than 4–5) cells should be used if characteristics of primary, somatic
cells wish to be examined.
ES Cell Culture
For detailed protocols on ES cell culture, readers are urged to consult
dedicated source of information, Hogan et al.20 and Robertson.21 For
proper maintenance of ES cell lines in their undifferentiated, pluripotent
state, a well-equipped tissue culture facility and rigorous culture conditions
are required. Some general recommendations include the use of sterile, dis-
posable plasticware or glassware that has never been exposed to detergent
and is kept separate from general laboratory supplies, as ES cells are highly
sensitive to trace levels of detergent. Only high-quality water, classified as
‘‘Type I Reagent Grade Water’’ (ASTM standard), should be used. Fetal
bovine serum should be purchased as ‘‘ES cell grade’’ or various batches
must be tested in parallel for quality control. Tissue culture grade plastic
flasks or plates, or glass coverslips (ESCO, cat. no. 9611301) are used for
ES cell culture and in all cases they have to be first coated in filter-sterilized
0.1% gelatin (Type A, Sigma cat. no. G-2500) in PBS 1X (PBS 10X, GIB-
CO, cat. no. 70013–016; sterile water, Invitrogen, 15230–089). The gelatin
solution is left in contact with the plastic for a minimum of 30 min
(it should not be allowed to dry out) and should be aspirated just prior to
addition of cell medium. We grow ES cells at 37 in 8% CO2.
ES cells can be maintained in the undifferentiated state either through
culture on mitotically inactivated feeder cells (such as mouse embryonic
fibroblasts) and/or in the presence of leukemia inhibitory factor (LIF),
depending on the cell line used. It should be noted that, for X inactivation
studies, ES cells should prefererably be cultured on male rather than
female feeder cells in order to avoid interference during analysis due to
the presence of the inactive X chromosome in female fibroblast cells. For
X inactivation kinetics studies, we have mainly used two female ES cell
lines, LF2 (a gift from Dr. Austin Smith) and PGK12.1 (a gift from Dr. Neil
Brockdorff), both of which grow on gelatin-coated flasks or plates,22 with-
out the need for feeder cells. It should be noted that similar X inactivation
kinetics have been found using feeder-dependent ES cell lines. ES cells are
maintained as an undifferentiated culture in DMEM with GlutMAX, 15%
fetal calf serum (GIBCO, cat. no. 16141–079), 104 mM 2-mercaptoethanol
21
E. J. Robertson, in ‘‘Teratocarcinomas and Embryonic Stem Cells: A Practical Approach’’
(E. J. Robertson ed.), p. 71. IRL Press, Oxford, England, 1987.
22
A. G. Smith, J. Tiss. Cult. Meth. 13, 89 (1991).
410 immunochemical assays of chromatin functions [27]
(Sigma, cat. no. M7522), and 1000 U/ml LIF (Chemicon, cat. no.
ESG1107). ES cells should be plated and maintained at relatively high
density and passaged at 70–80% confluence. This usually means passaging
every couple of days (cell doubling time can be between 8 and 22 h
depending on the ES cell line and serum used). ES cell colonies should
be monitored daily for density and also any signs of differentiation
(see refs. 20 and 21 for feeder cell–dependent ES cells and Smith22 for
non–feeder-dependent cell lines). The medium should be changed 3 h
prior to trypsinization to increase viability of the cells upon passaging or
cryopreservation. To passage, the flask or plate is rinsed with PBS, then
trypsin-EDTA is added (GIBCO, cat. no. 25200–072). After 8 min in the
incubator, an equal volume of ES cell medium is added and the cells are
dispersed into a single cell suspension by vigorous pipetting (20–40 times)
with a plugged Pasteur pipette, then centrifuged. This resuspension step
is important, as undissociated clumps of ES cells will rapidly form large
colonies and begin to differentiate before the next passage. Cells should
then be plated at 2–4 104 cells/cm2. For freezing, 1 106 cells
should be resuspended in 1 ml of FCS/10% DMSO, then put at 80 O/
N then into liquid nitrogen. The number of passages should always
be kept to a minimum for ES cells in order to avoid genetic and
karyotypic abnormalities. If possible, multiple vials should be frozen at
every passage.
ES Cell Differentiation
Differentiation of ES cells can be induced using a variety of strategies.
The most classical method involves differentiation into embryoid bodies
(EBs) by removal of feeder cells,21 LIF withdrawal, and culture of aggre-
gated (lightly trypsinized) cells in non-adherent Petri dishes (plastic dishes
used for bacterial agar plates are ideal) in DMEM supplemented with 10%
FBS. EBs have to be cultured in suspension, at least for the first 4 days,
which is the period during which X inactivation is established. In order
to perform immunofluorescence and/or RNA FISH on EBs, cytocentrifu-
gation is required, which can disrupt interphase chromatin structure and
nuclear architecture. An alternative strategy (based on ref. 22), which we
prefer as it allows cells to be grown directly on coverslips during differ-
entiation, involves the use of retinoic acid. ES cells are plated at a low
density of 104 cells/cm2 (if necessary after removal of feeder cells by ad-
sorption). Once the majority of cells have attached (this usually takes
8 h–overnight), the ES cell medium is removed, the cells are washed three
times in PBS to eliminate LIF, and differentiation medium is added. Dif-
ferentiation medium is DMEM supplemented with 10% FBS, 100 nM
[27] X-chromosome inactivation in mouse embryonic stem cells 411
all-trans-retinoic acid (RA) (Sigma, cat. no. R2625; a 103 M stock solution
is prepared in ethanol and stored at 20 ), and 104 mM 2-mercaptoetha-
nol. Addition of the latter has been found to minimize cell death. Differen-
tiation medium is changed daily and can be preceded by one or two washes
in PBS to remove debris if necessary. Changes in cell morphology can
be detected by day 2, with clearly fibroblast-like and neuronal-like cells
by day 4.
Immunofluorescence
Sub-confluent fibroblasts or ES cells cultured on gelatin-coated cover-
slips are briefly washed once in PBS and then fixed in freshly prepared
3% formaldehyde (paraformaldehyde, RE/PURO cat. no. 387507, in
PBS, pH 7.2) for 10 min at RT. Following two washes in PBS at RT,
23
D. L. Spector, R. D. Goldman, and L. A. Leinwand, ‘‘Cells: A Laboratory Manual.’’ Cold
Spring Harbor Laboratory Press, Cold Spring Harbor, NY, 1998.
24
R. H. Singer, J. B. Lawrence, and C. Villnave, Biotechniques 4, 230 (1986).
25
A. M. Femino, F. S. Fay, K. Fogarty, and R. H. Singer, Science 280, 585 (1998).
412 immunochemical assays of chromatin functions [27]
26
E. Harlow and D. Lane, ‘‘Antibodies: A Laboratory Manual.’’ Cold Spring Harbour
Laboratory Press, Cold Spring Harbor, NY, 1989.
[27] X-chromosome inactivation in mouse embryonic stem cells 413
TABLE I
Enriched (þ)
depleted ()
Antibody on the Xi Dilution Source
a
Note: The reader is referred to the chapter by Samra and Reinberg for characterization of
the anti-H3 di-/tri-methylated Lys-27 antibody.
Immunofluorescence
The technique we describe involves IF performed on unfixed meta-
phase chromosomes and is based on a previously published protocol of
ref 13. Metaphase chromosomes are prepared from cells at 80% conflu-
ence, following a 1-h incubation with 0.1 g/ml colcemid (Roche,
295892). To maximize the collection of mitotic cells the supernatant is col-
lected and adherent cells are trypsinized, pelleted, resuspended in PBS, and
pelleted (this wash performed twice). The cells are then resuspended (105
cells/ml) carefully in 75 mM KCl, incubated for 10 min at RT and then
placed on ice for between 5 min and 1 h. Cytocentrifugation is then per-
formed (Shandon), with the time and speed of centrifugation depending
on the cells used. For ES cells and early stages of differentiation, 8 min
at 1200 rpm is usually appropriate. After air-drying, the slides are incu-
bated in freshly prepared KCM buffer (120 mM KCl, 20 mM NaCl,
10 mM Tris-HCl, pH 8, 0.5 M EDTA, 0.1% Triton X-100) for 15 min at
RT, and incubated successively with primary and secondary antibody in
KCM, 1% BSA for 40 min in a humidified dark chamber at RT (immuno-
fluorescence protocol as earlier). Several washes in KCM buffer are per-
formed. A postfixation step in 3% formaldehyde can be performed at this
stage to prevent signal loss if this is a problem. The slides should again be
washed in KCM containing DAPI (0.2 mg/ml DAPI) for 2 min, washed
once more in KCM and mounted in medium (see earlier) for visualization
on a fluorescence microscope. An example is shown in Fig. 2.
[27]
[27] X-chromosome inactivation in mouse embryonic stem cells 417
be removed using forceps before the second wash). The samples should
then be dehydrated through a series of 3-min washes in 70%, 80%, and
95% ethanol. Denaturation of the DNA is performed by plunging the slide
in a 50-ml falcon tube or coplin jar containing 70% formamide/2 SSC (pH
7.2) for 3 min at 75 . The slides should immediately be plunged into ice-
cold 70% ethanol and then dehydrated (as earlier). Denaturation of
chromosome paint probes and pre-annealing of competitor DNA as well
as in situ hybridization are performed exactly as recommended by the sup-
plier (mouse X-chromosome paint, Cambio, cat. no. 1187-XMB-02). Hy-
bridization is performed overnight at 42 . Posthybridization washes are at
45 , three times in 50% formamide/2 SSC and three times in 2 SSC.
Slides are counter-stained with DAPI, mounted, and viewed under the
fluorescence microscope (as earlier). An example is shown in Fig. 2.
27
B. Dutrillaux, C. Laurent, J. Couturier, and J. Lejeune, Acad. Sc. Paris 276, 3179 (1973).
28
N. Takagi and M. Oshimura, Exp. Cell. Res. 78, 127 (1973).
418 immunochemical assays of chromatin functions [27]
the tubes are kept on ice. Clean and dry slides are placed on wet tissues
(Kimwipes, Kimberly-Clark). The fixed cell suspension is spread by drop-
ping one drop onto an ethanol-cleaned slide from a height of about
10 cm. The slides are allowed to dry very quickly. They are then stained
with Sorensen’s phosphate buffer (pH 6.8) supplemented with 0.05 mg/ml
acridine orange solution. The slides are incubated for 10–15 min in this so-
lution, rinsed immediately in running tap water for 1–2 s, then rinsed in
phosphate buffer (pH 6.8). They are mounted with a coverslip in phosphate
buffer (pH 6.8) and excess buffer is removed with blotting paper. The
metaphase spreads are then immediately examined under a fluorescence
microscope4,27,28 using a filter block for acridine orange (blue filter, exc.
490 nm, emiss. 530 nm/640 nm for single- and double-stranded DNA
detection). Whole chromosomes or chromosome segments replicated in
the presence of BrdU show a weak red fluorescence whereas, those that
have not incorporated BrdU show bright green fluorescence. The auto-
somes and the active, synchronously replicating X chromosome have
certain regions that replicate late in the S phase of the cell cycle as well
as others that replicate early. When BrdU is present only during the latter
half of S phase, an unequivocal red and green banding pattern is observed
in these chromosomes in contrast with the late-replicating inactive
X chromosome, which is stained homogeneously red. An example is shown
in Fig. 3.
Conclusion
The techniques we describe here should provide a basis for investigat-
ing the onset of X inactivation or other developmentally regulated pro-
cesses in differentiating ES cells. Using these techniques we have been
able to determine the relative order of the following events: Xist RNA
coating occurs within the first 24–48 hours of differentiation; histone H3
modifications such as hypomethylation of Lys-4, hypoacetylation of Lys-
9, and hypermethylation of Lys-9 and Lys-27 (see Fig. 1) are detectable
on the X chromosome in a proportion of interphase cells as soon as Xist
RNA accumulates. Histone H4 hypoacetylation of lysines 5, 8, 12, and 16
is also observed from this time, although with slightly slower kinetics.15,17
The appearance of these marks on the X chromosome during mitosis (see
Fig. 2) is found from days 2–3, suggesting that they represent true epige-
netic marks. On the other hand, a late-replicating X chromosome is only
detected from day 3.17 Thus, histone modifications appear to be direct con-
sequences of Xist RNA’s action, while late replication timing seems to be a
slightly later event. Using this as a basis, the relationship between these dif-
ferent events can now be defined and tested, using for example, genetically
modified ES cells or siRNA-induced silencing of candidate modifying
enzymes.
1
K. Plath, S. K. Mlynarczyk-Evans, D. A. Nusinow, and B. Panning, Annu. Rev. Genet. 36,
233 (2002).
Conclusion
The techniques we describe here should provide a basis for investigat-
ing the onset of X inactivation or other developmentally regulated pro-
cesses in differentiating ES cells. Using these techniques we have been
able to determine the relative order of the following events: Xist RNA
coating occurs within the first 24–48 hours of differentiation; histone H3
modifications such as hypomethylation of Lys-4, hypoacetylation of Lys-
9, and hypermethylation of Lys-9 and Lys-27 (see Fig. 1) are detectable
on the X chromosome in a proportion of interphase cells as soon as Xist
RNA accumulates. Histone H4 hypoacetylation of lysines 5, 8, 12, and 16
is also observed from this time, although with slightly slower kinetics.15,17
The appearance of these marks on the X chromosome during mitosis (see
Fig. 2) is found from days 2–3, suggesting that they represent true epige-
netic marks. On the other hand, a late-replicating X chromosome is only
detected from day 3.17 Thus, histone modifications appear to be direct con-
sequences of Xist RNA’s action, while late replication timing seems to be a
slightly later event. Using this as a basis, the relationship between these dif-
ferent events can now be defined and tested, using for example, genetically
modified ES cells or siRNA-induced silencing of candidate modifying
enzymes.
1
K. Plath, S. K. Mlynarczyk-Evans, D. A. Nusinow, and B. Panning, Annu. Rev. Genet. 36,
233 (2002).
2
G. D. Penny, G. F. Kay, S. A. Sheardown, S. Rastan, and N. Brockdorff, Nature 379, 131
(1996).
3
Y. Marahrens, B. Panning, J. Dausman, W. Strauss, and R. Jaenisch, Genes Dev. 11, 156
(1997).
4
G. Borsani, R. Tonlorenzi, M. C. Simmler, L. Dandolo, D. Arnaud, V. Capra, M. Grompe,
A. Pizzuti, D. Muzny, C. Lawrence et al., Nature 351, 325 (1991).
5
N. Brockdorff, A. Ashworth, G. F. Kay, P. Cooper, S. Smith, V. M. McCabe, D. P. Norris,
G. D. Penny, D. Patel, and S. Rastan, Nature 351, 329 (1991).
6
C. J. Brown, A. Ballabio, J. L. Rupert, R. G. Lafreniere, M. Grompe, R. Tonlorenzi, and
H. F. Willard, Nature 349, 38 (1991).
7
C. J. Brown and S. E. Baldry, Somat. Cell Mol. Genet. 22, 403 (1996).
8
C. M. Clemson, J. C. Chow, C. J. Brown, and J. B. Lawrence, J. Cell Biol. 142, 13 (1996).
9
C. M. Clemson, J. A. McNeil, H. F. Willard, and J. B. Lawrence, J. Cell Biol. 132, 259
(1996).
10
B. Panning and R. Jaenisch, Genes Dev. 10, 1991 (1996).
11
B. Panning, J. Dausman, and R. Jaenisch, Cell 90, 907 (1997).
12
S. A. Sheardown, S. M. Duthie, C. M. Johnston, A. E. Newall, E. J. Formstone, R. M.
Arkell, T. B. Nesterova, G. C. Alghisi, S. Rastan, and N. Brockdorff, Cell 91, 99 (1997).
13
A. Wutz and R. Jaenisch, Mol. Cell 5, 695 (2000).
14
S. F. Wolf, D. J. Jolly, K. D. Lunnen, T. Friedmann, and B. R. Migeon, Proc. Natl. Acad.
Sci. USA 81, 2806 (1984).
[28] X inactivation in mouse ES cells 421
Differentiation of ES Cells
For X inactivation studies, differentiation of ES cells into embryoid
bodies (EBs) using hanging drop cultures is very effective. When the
resulting EBs are placed on a tissue culture substrate coated slides the
EBs attach to the slides. The cells that delaminate and migrate away from
the EBs can be analyzed at various intervals after plating to visualize
the alterations in chromatin structure that occur during X chromosome
silencing.
1. Trysinize and pellet female ES cells.
2. Resuspend the cell pellet in 1 ml of differentiation media.
Differentiation media:
DMEM
15% fetal bovine serum
Non-essential amino acids
0.1 mM -mercaptoethanol
15
P. Jeppesen and B. M. Turner, Cell 74, 281 (1993).
16
E. Heard, C. Rougeulle, D. Arnaud, P. Avner, C. D. Allis, and D. L. Spector, Cell 107, 727
(2001).
17
J. E. Mermoud, B. Popova, A. H. Peters, T. Jenuwein, and N. Brockdorff, Curr. Biol. 12,
247 (2002).
18
J. Silva, W. Mak, I. Zvetkova, R. Appanah, T. B. Nesterova, Z. Webster, A. H. Peters,
T. Jenuwein, A. P. Otte, and N. Brockdorff, Dev. Cell 4, 481 (2003).
19
K. Plath, J. Fang, S. K. Mlynarczyk-Evans, R. Cao, K. A. Worringer, H. Wang, C. C.
de la Cruz, A. P. Otte, B. Panning, and Y. Zhang, Science 300, 131 (2003).
20
C. Costanzi and J. R. Pehrson, Nature 393, 599 (1998).
21
N. Takagi, Exp. Cell Res. 86, 127 (1974).
22
M. L. Barr, C. D., Acta Cytol. 6, 34 (1961).
422
immunochemical assays of chromatin functions
Fig. 1. Immunostaining for Eed and FISH for Xist RNA in differentiating female ES cells. In a differentiating population of female ES
cells, a fraction of cells exhibit Xist RNA coating of the Xi (middle panel) in nuclei, visualized DAPI staining (left panel). In these cells Eed,
a component of a Polycomb Group complex that includes the histone methyltransferase Ezh2, is also enriched on the Xi (right panel).
Recruitment of the Eed-Ezh2 complex correlates with an increase in histone H3 lysine 27 methylation on the Xi.18,19 In the cells that do not
yet show Xist RNA coating, the levels of Eed are much higher, suggesting that abundance of Eed decreases when ES cells are differentiated.
[28]
[28] X inactivation in mouse ES cells 423
Immunostaining
To simultaneously visualize protein and RNA, immunostaining is
followed by FISH. Many antisera contain RNases, resulting in loss of RNA
during the immunostaining procedure. Addition of tRNA and RNase in-
hibitors to the immunostaining buffers to suppress the RNases in the sera
preserves RNA during immunostaining.
1. Rinse slides with 1 PBS.
2. Incubate slides for 30 s in ice-cold cytoskeletal buffer.
Cytoskeletal buffer:
100 mM NaCl
300 mM sucrose
3 mM MgCl2
10 mM PIPES, pH 6.8
3. Incubate slides for 30 s in ice-cold cytoskeletal buffer þ 0.5%
Triton X-100.
4. Incubate slides for 30 s in ice-cold cytoskeletal buffer.
5. Fix in 4% paraformaldehyde/1 PBS, 10 min, room temperature.
6. Wash slides two times, 5 min each in room temperature PBS-
Tween. Slides can be stored for several days in PBS-Tween at 4
prior to immunostaining.
424 immunochemical assays of chromatin functions [28]
PBS/Tween
1 PBS
0.2% Tween 20
7. Block slides for 30 min at 37 in blocking buffer in a humid
chamber. The humid chamber should contain PBS/Tween.
Blocking buffer:
1 PBS
5% goat serum (heat inactivated)
0.2% Tween 20
0.2% fish skin gelatin
4–5 units/ml RNAsin
1 mg/ml molecular biology-grade yeast tRNA
8. Incubate with primary antibody diluted in blocking buffer, 1 h, 37
in a humid chamber. Temperature may be varied depending on the
primary antibody.
9. Wash slides three times, 5 min each with PBS/Tween.
10. Block slides for 5 min at 37 in blocking buffer in humid chamber.
11. Incubate with secondary antibody diluted in blocking buffer for
30 min, 37 in a humid chamber.
12. Wash slides three times, 5 min each with PBS/Tween. Minimize
exposure to light to limit photobleaching.
13. Fix in 2% paraformaldehyde/1 PB for 10 min at room tempera-
ture.
14. Transfer slide to 70% ethanol. Slides can be stored in 70% ethanol
at 4 for several weeks prior to carrying out FISH.
FISH
The FISH procedure involves generating a labeled probe, hybridizing
the probe to the fixed and immunostained sample, and then detection
of the labeled probe. The specific activity of RNA probes is highest, and
since the immunostaining may cause a slight reduction in the amount of
RNA in the sample, using high specific activity probe generates the best
results with FISH. All stock solutions are autoclaved to minimize likeli-
hood of RNase contamination, and all stock solutions are diluted in auto-
claved water.
Generate Template
Templates can be generated by cloning fragments into plasmids with
T3, T7, or SP6 promoters or by appending T3, T7, or SP6 promoters to
PCR products. The maximum size of an in vitro transcribed RNA that will
[28] X inactivation in mouse ES cells 425
Prepare Template
If using plasmid template, digest with appropriate restriction enzyme to
linearize. Phenol-chloroform extract, ethanol precipitate twice, and dry.
Resuspend in water at a concentration of approximately 1 mg/ml.
If using PCR product, set up PCR reaction and run fraction out on gel.
If the band is clearly visible, then about 1 l of PCR product is enough for the
in vitro transcription reaction. It does not need any further manipulation.
In Vitro Transcription
1. Mix the following at room temperature:
5 buffer 20 l
100 mM DTT 10 l
4000 units/ml RNA sin 2.5 l
2.5 mM A, G, UTP mix 20 l
10 mM CTP 1 l
10 mM bioCTP 4 l
DNA, 1 mg/ml 2 l
Polymerase 2 l
Water 38.5 l
2. Incubate at 37 for 1 h.
3. Add 1 l of 1 U/l RNase free DNase 1.
4. Remove unincorporated nucleotides using G-50 Sephadex spun
columns, prepared with TE.
5. After RNA has spun through column, run 5 l on agarose minigel to
ensure that in vitro transcription has worked. The remaining RNA
should be ethanol precipitated with 10 g of molecular biology-
grade yeast tRNA and stored at 80 or 20 under ethanol.
FISH
1. Dehydrate cells on slides through an ethanol series, 2 min each in
85%, 95%, and 100% ethanol, air-dry.
2. Add probe to air-dried cells.
3. Place coverslip over probe. For a 12-mm circle coverslip use 3 l of
probe; for a 22-mm square coverslip use 10 l.
4. Incubate in a humid chamber at 37 overnight. The humid chamber
should contain 2 SSC/50% formamide.
Troubleshooting
Generally it is best to test the FISH probe on fixed material prior to
doing immunostaining in combination with FISH. Follow the first five fix-
ation steps of the immunostaining protocol. After the paraformaldehyde
fixation, transfer slides to 70% ethanol. Follow the FISH protocol. If the
428 immunochemical assays of chromatin functions [28]
FISH does not work, then troubleshooting of the probe and solutions is
necessary. Introduction of RNases is one likely cause of failure, and using
freshly prepared, autoclaved solutions can solve this problem. The amount
of labeled probe that is best for detection of different RNAs varies and
testing a range of concentrations of labeled probe can be used to identify
an optimal probe concentration.
RNA has been implicated in regulating chromatin structure at centro-
meres23–25 and in recruiting chromatin modifying enzymes to the dosage
compensated X chromosome in Drosophila,26,27 indicating that RNA
may be more generally used to regulate chromatin structure. Thus, the
techniques described in this chapter may be relevant to the study of other
examples of regulated changes in chromatin structure.
Acknowledgments
Many thanks to Hannah Cohen, Susanna Mlynarczyk-Evans, and Kathrin Plath for
critical reading of this manuscript.
23
C. Maison, D. Bailly, A. H. Peters, J. P. Quivy, D. Roche, A. Taddei, M. Lachner,
T. Jenuwein, and G. Almouzni, Nature Genetics 30, 329 (2002).
24
T. A. Volpe, C. Kidner, I. M. Hall, G. Teng, S. I. Grewal, and R. A. Martienssen, Science
297, 1833 (2002).
25
I. M. Hall, G. D. Shankaranarayana, K. Noma, N. Ayoub, A. Cohen, and S. I. Grewal,
Science 297, 2232 (2002).
26
V. H. Meller, K. H. Wu, G. Roman, M. I. Kuroda, and R. L. Davis, Cell 88, 445 (1997).
27
H. Amrein and R. Axel, Cell 88, 459 (1997).
Author Index
Numbers in parentheses are footnote reference numbers and indicate that an author’s work is
referred to although the name is not cited in the text.
429
430 author index
Bottcher, B., 60 C
Bouillon, R., 343
Bouman, P., 314, 349, 357(1), 360(1) Caiafa, P., 76
Bourachot, B., 120 Cairns, B. R., 120, 317
Boyer, L. A., 106(12), 107 Caldwell, C. W., 333
Brachmann, C. B., 172(14), 173 Caldwell, J. A., 231
Brady, J. N., 196 Callaghan, J., 132, 137(10), 140(10), 146(10),
Brame, C. J., 231 147, 151, 161(15)
Brandts, J. F., 146, 147(40) Cam, H., 306, 318, 360
Branton, D., 32 Campbell, M. J., 360
Brasher, S. V., 149, 151(7), 167(7) Cannova, C., 174
Braunstein, M., 349 Cantor, R. C., 140, 142(30)
Breslow, R., 171, 199, 200(6) Cao, R., 248, 252, 254(28; 29), 260, 262, 332,
Briggs, S. D., 131, 132(8), 134(8), 140(8), 405, 406(10), 408(10), 421, 422(19)
142(8), 144(8), 231 Capra, V., 421
Brinkley, B. R., 232 Carlson, J. E., 121, 125(20), 126, 127(20), 128,
Broach, J. R., 349 128(20), 129(53), 130(20; 53), 137, 233
Broadhurst, R. W., 149, 151(6; 7), 167(7), Carlson, M., 289
168(6), 169(6) Carmen, A. A., 209, 217(7), 289, 293(1)
Brockdorff, N., 241, 242(24), 247(24), Caron, C., 175
248, 249(30), 254(30), 260, 405, Caron, F., 92
406, 406(9), 408(9; 16), 420, Carozza, M. J., 120
421, 422(18) Carroll, A. S., 357
Brosch, G., 174, 176 Carruthers, L. M., 17, 20(7; 8)
Brower-Toland, B., 62, 65, 65(8), 67(8), Cavagnero, S., 125
68(8), 71(8) Cavalli, G., 374, 392
Brown, C. E., 120 Cavanagh, J., 165(31), 166
Brown, C. J., 420 Cawley, S. E., 317
Brown, D. R., 199(8), 200 Cebrat, M., 188, 191, 193(10), 194(10)
Brown, P. O., 290, 302, 314, 317, 323(12), 350, Cechvala, M., 318
353, 353(10), 354(10), 357(13), 360 Celic, I., 172(14), 173, 174
Brownell, J. E., 106, 117(29), 119, 120, 189, Chadwick, B. P., 405
199, 223 Chakraburtty, K., 290
Brownstein, M. J., 323 Chalkley, R., 336
Brunger, A. T., 124, 139 Chambon, P., 119, 120(4), 121(4), 127(4)
Buchard, J., 304 Chang, J. H., 173
Bulger, M., 65(16), 66 Chang, J. J., 60
Burakoff, S. J., 123 Charlton, J. A., 328
Burgoyne, P. S., 237, 242(22), 246(22), Chase, M. W., 380
248(22), 254(22) Chatenay, D., 92
Burlingame, R. W., 217 Chaumeil, J., 405, 406, 408(17), 419(17)
Burnett, E., 316 Chaya, D., 361, 368(3)
Burns, J. A., 164 Chehab, N. H., 128
Bustamante, C., 62, 76, 82, 92 Chen, C. M., 328, 333, 333(30)
Bustin, M., 112, 209, 211, 212(3), 213(9), 214, Chen, D., 173, 273
214(3), 215, 216(15), 361, 368(3) Chen, H., 45
Butler, J. C., 164 Chen, M., 323
Buttle, K., 44 Chen, T., 314(9), 315, 317
Byrd, R. A., 149 Chen, Y., 323
Bystricky, K., 17 Cheng, A., 60
432 author index
Cheung, P., 171, 221, 223, 232(15), 233(5), Copeland, R. A., 187
237, 238(21), 247(21), 405 Cornish-Bowden, A., 187
Cheung, W. L., 223, 232(15) Corona, D. F., 12
Chin, J., 3 Correira, J. J., 164
Chinault, A. C., 222, 237, 238(21), Costanzi, C., 405, 421
247(21), 405 Costanzo, A., 197
Chinnaiyan, A. M., 270 Cote, J., 235
Chiu, W., 39, 41, 42(12) Cotter, R. J., 193, 194(14)
Cho, R. J., 360 Courtney, A. D., 193, 194(14)
Cho, Y., 173 Couturier, J., 417
Choi, I., 316 Cowan, S. W., 59
Chou, J. J., 125 Cowell, I. G., 237, 242(22), 246(22),
Chow, J. C., 420 248(22), 254(22)
Chowdhry, B. Z., 147 Cowieson, N. P., 139, 149, 151(8)
Christensen, J., 316 Coyne, R. S., 140, 142(29)
Christopherson, R. I., 163 Craig, C. A., 372, 373, 386, 386(9), 387,
Chu, S., 74, 80(4) 387(8), 389(9), 394, 403
Chua, P., 120 Crane-Robinson, C., 222, 336
Chuikov, S., 248, 254(27), 255 Creech, H. J., 379(16; 17), 380
Chung, W. H., 49, 59(4) Croce, K., 139
Church, G. M., 360 Croquette, V., 76, 101(6), 102(5)
Churchill, M. E. A., 108, 117(17) Cross, S. H., 328
Ciccone, C. N., 334 Crowley, K. A., 17, 106(12), 107
Cimato, T. R., 272 Crowther, R. A., 45, 47(21), 60
Clayton, A. L., 219, 222, 223, 232(16) Crumley, T. M., 174
Clayton, D. F., 361 Cui, Y., 62, 82, 92
Cleary, M., 283 Czermin, B., 252, 262
Clements, A., 108, 109(14), 111(14), 112(14)
Clemson, C. M., 420
D
Clore, G. M., 123, 124, 136, 139
Cluzel, P., 92 Dahlquist, F. W., 166
Cocito, A., 315 Dai, H., 290, 304, 313, 359
Coessens, B., 315 Damelin, M., 317
Coffey, E., 290 Dandolo, L., 420
Cohen, A., 171, 428 Daniel, C., 3
Cohen, H. Y., 185 Dannenberg, A. J., 195
Cohen, L. H., 372, 374, 374(2) Danzer, J. R., 390
Cole, P. A., 114, 115, 188, 189, 190, 191, Darkin-Rattray, S. J., 174
191(7), 192, 193, 193(7; 10; 13), 194(9; 10; Darnell, J. E., Jr., 361
14), 195, 196, 197, 197(12), 198 Daugherty, M., 191
Cole, R. N., 174 Daujat, S., 272, 283(15; 16)
Colletti, S. L., 174 Dausman, J., 420
Compton, J. L., 404 Davidson, I., 119, 196
Connors, B., 222 Davie, J. R., 174, 346
Constantini, F., 408, 409(20) Davis, R. L., 428
Conway, J. F., 42 Davuluri, R. V., 315
Cook, R. G., 106(13), 107, 117(29; 30), 119, Dawid, I. B., 120
222, 223, 223(9), 231 De Angelis, J., 191, 194(9)
Coons, A. H., 379(16; 17), 380 Decallonne, B., 343
Cooper, P., 420 de Grooth, B. G., 62, 93, 94, 95(29), 96(29)
author index 433
de la Cruz, C. C., 248, 254(29), 260, 405, Dynlacht, B. D., 304, 306, 318, 361
406(10), 408(10), 421, 422(19) Dyson, H. J., 125, 130(44), 137
Delaglio, F., 123, 136 Dyson, M. H., 121
DeLano, W. L., 139 Dziedzic, J. M., 74, 80(4)
Delcuve, G. P., 174
Demeler, B., 18, 19(12), 20(12), 24(12)
E
Demeter, J., 359
De Moor, B., 315 Easom, R. A., 18, 27(11)
Deng, H., 173 Eberharter, A., 106(13), 107
Denu, J. M., 106(12), 107, 114, 131, 173, 174, Ebert, A., 373, 387(10), 389(10), 392(10)
177(21; 24), 178, 194, 223, 232(15) Ebright, R. H., 140
DeRisi, J. L., 314, 350, 357 Eden, A., 408
DeRosier, D. J., 45, 47(21) Edmondson, D. G., 106, 117(29), 119, 120,
Dessain, S. K., 173 189, 199, 223, 267
Deuring, R., 120 Eftink, M. R., 157(21), 158, 160(21)
Devlin, M. K., 188 Eick, D., 316
Dhalluin, C., 121, 125(20), 126, 127(20), Eiletici, P., 125
128(20), 130(20), 137, 233 Einck, L., 30, 215
Dhanasekaran, S. M., 270 Eisen, M. B., 302, 360
Diaz, M. O., 295, 353 Eisenberg, H., 5, 255
Dickens, N. J., 147 Eisenhaber, F., 147, 287, 387
Dietrich, V., 373, 386, 386(9), 389(9), 394 Eissenberg, J. C., 131, 132(8), 134(8), 140(8),
Dignam, J. D., 20 142(8), 144(8), 147, 151, 373, 386, 386(9),
Dimitrov, S., 18 389(9), 394
Dollard, C., 120 Ekstrom, T. J., 171
Domaille, P. J., 149, 151(6), 168(6), 169(6) Ekwall, K., 392
Dombrowski, A. W., 174 Elgin, S. C. R., 372, 373, 374, 376(1), 380,
Dong, H., 350 380(1), 386, 387, 387(8), 389(9), 394, 403
Donigian, J. R., 171 Ellis, C. D., 269
Dorigo, B., 17 Ellis, T. W., 272
Doring, P., 395 Emre, N. C., 237, 270, 271(7), 283(7)
Dorn, R., 373, 387, 387(10), 389(10), 392(10) Erard, M., 316
Dorr, A., 130, 197 Erdjument-Bromage, H., 120, 231, 235, 248,
Douarin, B. L., 119, 120(4), 121(4), 127(4) 252, 254(27; 28), 255, 262, 332, 395, 408
Dow, S., 304 Erlich, R. L., 314, 349, 357(1), 360(1)
Downing, K. H., 45 Escalante-Semerena, J. C., 172(14), 173, 174
Doyle, M., 237, 238(20), 240(20), 241(20), Espinosa, R. III, 295, 353
242(20), 246(20), 247(20) Ettinger, M. J., 272
Drenkow, J., 317 Euskirchen, G., 317
Driscoll, P. C., 136 Evans, P. R., 125, 137
Du, J., 120 Ezeokonkwo, C., 48
Du, K., 196
Dubochet, J., 60
F
Duggan, L. J., 120
Dulski, P. M., 174 Fairbrother, W. J., 165(31), 166
Dunu, J. M., 172(16; 17), 173, 177(16) Fanciulli, M., 197
Duquet, A., 192, 197(12) Fang, J., 248, 254(28; 29), 260, 405, 406(10),
Duthie, S. M., 420 408(10), 421, 422(19)
Dutnall, R. N., 112 Fanti, L., 237, 242(22), 246(22),
Dutrillaux, B., 417 248(22), 254(22)
434 author index
Gordon, D. B., 314, 317 Hannett, N., 290, 314, 317, 322(13), 333(13),
Gotchel, B. V., 374 350, 353(11), 357(11)
Gottschling, D. E., 235 Hansen, J. C., 17, 18, 18(5; 6), 19(5; 6; 12),
Gounalaki, N., 120 20, 20(5–9; 12), 22, 24(12), 26(6),
Grant, P. A., 106, 106(11; 13), 107, 120 27(5; 6), 33
Grassucci, R. A., 60 Hansen, U., 272
Graveel, C. R., 332 Harbison, C. T., 314, 317
Gray, S. G., 171 Harel-Bellan, A., 192, 197(12)
Green, G. R., 30 Harkiolaki, M., 139
Greene, J., 315 Harlan, J. E., 123
Greenleaf, A. L., 372, 374(3), 394 Harlow, E., 212, 216(11), 228, 368, 412
Greenwald, M., 200 Harnett, T., 151
Greve, J., 62, 93, 94, 95(29), 96(29) Harper, S., 112, 114(23)
Grewal, S. I. S., 151, 335, 388, 428 Harrelson, A. L., 361
Griesinger, C., 128 Harris, K. G., 128
Griess, G. A., 18, 27(11) Harrison, P. M., 317
Grigliatti, T. A., 386 Hart, C. M., 252, 262
Grompe, M., 420 Hartman, S., 317
Gronenborn, A. M., 123, 124, 136 Hartnett, T., 386
Gros, P., 139 Hassig, C. A., 174, 189, 199
Grosse, I., 315 Haugland, R. P., 142
Grosse-Kunstleve, R. W., 139 Hayamizu, T., 361, 368(3)
Grozinger, C. M., 171, 172(8) Hayes, J. J., 171
Grubmeyer, C., 172(14), 173 Haynes, S. R., 120
Grummt, I., 174 He, C., 121, 125(20), 126, 127(20), 128(20),
Grunstein, M., 171, 209, 217(7), 289, 290, 130(20), 137, 233
290(3), 292, 293(1), 294, 294(9), 300, He, Y. D., 128, 129(53), 130(53), 290, 313, 359
300(10; 14), 301, 302(10), 303, Heagle, A. B., 60
304(14), 314(38), 315, 349, 351(7), Heard, E., 237, 238(21), 247, 247(21), 260,
352(7), 360(2) 405, 406, 408(15; 17), 419(15; 17), 421
Grzesiek, S., 123, 136, 137 Hebbes, T. R., 222, 336
Gu, W., 128, 173 Hebert, H., 36
Guarente, L., 120, 172(13), 173, 177(13) Hebert, J. M., 359
Guggiari, M., 406, 408(17), 419(17) Hecht, A., 290, 292, 300(14), 304(14), 349,
Gurnett, A. M., 174 351(7), 352(7)
Guzman, E., 395 Heller, C., 92
Heller, R. C., 172(15), 173, 175(15), 177(15),
182(15), 183(15)
H
Hendzel, M. J., 174, 232
Hager, G. L., 28, 196 Henklein, P., 130, 197
Hahn, M., 48 Hernandez-Boussard, T., 359
Hajduk, P. J., 127 Herrera, J. E., 112
Halazonetis, T. D., 128 Herrera, R. E., 283
Haley, R. H., 109, 111(19), 112(19), 114(19), Hess, D., 175
118(19), 193, 198 Hesse, J., 36
Hall, I. M., 428 Hetzer-Egger, C., 197
Halperin, T., 316 Heyduk, T., 140
Halvorson, H. R., 164 Hilfiker, A., 152
Hamilton, S. L., 41 Hilfiker-Kleiner, D., 152
Hampsey, M., 248, 254(26) Hill, J. J., 4
436 author index
Losson, R., 119, 120(4), 121(4), 127(4) 117(16), 118(19), 128, 131, 188, 191, 193,
Louis, J. M., 125 193(10), 194, 194(10), 198
Loven, M. A., 316 Marsh, J. L., 200
Lowary, P. T., 5, 12(7) Marshall, B. L., 173, 174, 177(24)
Lowden, P. A., 200 Martienssen, R. A., 428
Lowe, N., 125, 137 Martinez-Yamout, M. A., 125, 130(44), 137
Loyola, A., 235, 255 Marton, M. J., 290, 304, 313, 359
Lu, H., 196 Martone, R., 317
Lucchesi, J. C., 152 Marzilli, L. A., 193, 194(14)
Ludtke, S. J., 39, 42(12) Mason, P., 395
Luger, K., 58, 59(9), 64, 65, 171, 196 Mastrangelo, M. A., 408
Lum, P. Y., 290 Matese, J. C., 359
Lunnen, K. D., 420 Mathieu, C., 343
Luo, J., 173 Matson, J. C., 124
Luo, K., 300 Matthias, G., 175
Luscombe, N. M., 317 Matthias, P., 175
Lusser, A., 174 Maurer-Stroh, S., 147
Lyon, M. F., 405 McCabe, D., 252, 262
McCabe, V. M., 420
McCampbell, A., 200
M McCarthy, B. J., 404
Ma, H., 231, 272 McDowall, A. W., 60
Ma, Y., 140 McEwen, B. F., 45
MacArthur, M. W., 125 Mcintosh, L. P., 166
Mader, A. W., 58, 59(9) McKinsey, T. A., 198
Madore, S. J., 332 McLaughlin, P. J., 139, 149, 151(8)
Mahadevaiah, S. K., 237, 242(22), 246(22), McNeil, J. A., 420
248(22), 254(22) McRorie, D. K., 162, 165(25)
Mahadevan, L. C., 121, 219, 223, Meadows, R. P., 123, 127
229(18), 232(16) Means, G. E., 157(20), 158, 276
Mahal, A., 200 Mechtler, K., 131, 151, 170(12), 234, 237,
Mahandiram, D. R., 123 238(19; 20), 240(20), 241(20), 242(20),
Maison, C., 213, 240, 271, 428 245(19), 246(19; 20), 247(20), 249(19),
Mak, W., 248, 249(30), 254(30), 260, 405, 254(19), 255, 287, 386, 387,
406(9), 408(9), 421, 422(18) 387(29), 392(29)
Makhatadze, G. I., 146 Medrano, E. E., 198
Mancini, M. A., 232 Meinke, P. T., 174
Manley, J. L., 128 Meisterernst, M., 316
Mansfield, K., 128 Melfi, R., 252, 262
Mao, D. Y. L., 318 Meller, V. H., 428
Marahrens, Y., 420 Mello, C., 196, 197(25)
Marcus, G. A., 120 Mellor, J., 237, 270, 271(7), 283(7)
Marion, D., 136 Melnyk, P., 60
Marko, J. F., 69(18), 72, 95 Menager, M., 33(26), 34
Marko, M., 44 Merlo, P., 197
Marks, P. A., 171, 199, 200, 200(6) Mermoud, J. E., 241, 242(24), 247(24), 260,
Marmorstein, R., 106, 108, 109, 109(14–16), 405, 406, 408(16), 421
111(14–16; 19), 112, 112(14–16; 19), Merril, C. R., 25
113(21), 114, 114(15; 16; 19; 21; 23), Mery, J., 174
440 author index
Pimpinelli, S., 237, 242(22), 246(22), Ranalli, T., 106, 117(29), 119, 120, 189, 199,
248(22), 254(22) 223, 232
Piotto, M., 166(34; 35), 167 Rasband, W. S., 25
Pirrotta, V., 252, 262 Raschke, E. E., 316
Pise-Masison, C. A., 196 Rasmussen, T. P., 408
Pizzuti, A., 420 Rastan, S., 405, 420
Plagens, U., 394 Ravichandran, K. S., 123
Platero, J. S., 151 Rea, S., 131, 151, 170(12), 255, 287, 373, 386,
Plath, K., 248, 254(29), 260, 405, 406(10), 387, 387(10; 29), 389(10), 392(10; 29)
408(10), 419, 421, 422(19) Read, R. J., 139
Ploton, M., 33(26), 34 Rechsteiner, T. J., 64
Poelman, M., 200 Reddy, V., 60
Poirier, G. G., 235 Reilly, C. M., 200
Polach, K. J., 5 Reinberg, D., 106, 175, 221, 233(4), 235, 248,
Polesskaya, A., 192, 197(12) 248(11), 252, 254(11; 26; 27), 255, 262,
Polishook, J. D., 174 272, 408
Ponting, C. P., 147, 255, 287, 387 Ren, B., 290, 304, 306, 314, 317, 318, 322(13),
Pope, L. H., 93 333(13), 350, 353(11), 357(11), 360
Popova, B., 406, 408(16), 421 Ren, G., 60
Portanova, J. P., 217 Reuter, G., 373, 387, 387(10), 389,
Postnikov, Y. V., 215, 216(15) 389(10), 392(10)
Poulsen, F. M., 124 Reyes, J. C., 120
Poux, A. N., 191, 193(10), 194(10) Rice, J. C., 151, 235, 248, 248(11), 254(11; 27),
Prantera, G., 237, 242(22), 246(22), 255, 388
248(22), 254(22) Rice, L. M., 139
Pray-Grant, M. G., 106(11), 107 Richard-Foy, H., 28
Prestegard, J. H., 125 Richmond, R. K., 58, 59(9)
Preuss, U., 232 Richmond, T. J., 17, 58, 59(9), 64, 65, 171
Price, D. H., 395 Richon, V. M., 171, 199, 200, 200(6)
Privalov, P. L., 146 Riegel, A. T., 28
Prives, C., 128 Riester, D., 176
Pullner, A., 316 Rifkind, R. A., 171, 199, 200(6)
Riggs, A. D., 362
Rinaldi, N. J., 314, 317
Q
Rinn, J. L., 317
Quate, C. F., 73 Robert, C., 76
Quivy, J. P., 240, 271, 428 Robert, F., 254, 290, 314, 317, 322(13),
333(13), 350, 353(11), 357(11)
Roberts, C. J., 290, 304, 313, 359
R
Roberts, S. M., 120
Rackwitz, H.-R., 130 Robertson, E. J., 409, 410(21)
Radermacher, M., 44, 53, 60 Robin, P., 192, 197(12)
Radonovich, M., 196 Robinson, R. C., 209
Rafty, L. A., 173 Robyr, D., 289, 290, 294(9), 300(10),
Rahmatpanah, F., 328, 333, 333(30) 301, 302(10), 303, 314(8), 315,
Raine, A. R. C., 149, 151(6), 168(6), 169(6) 349, 360(2)
Ralston, G., 163, 164(26) Roche, D., 240, 271, 428
Ramakrishnan, V., 112 Rodrigo, I., 174
Raman, C. S., 146, 147(39) Rodriguez-Alfageme, C., 372, 374(2)
Ramjit, R., 97, 103(32) Roeder, G. S., 120
author index 443
Roeder, R. G., 20, 76, 78(17), 114, 128, 189, Scafe, C. S., 290, 317, 350, 353(10), 354(10)
191(7), 193(7), 195, 196(20) Schalch, T., 17
Rohrer, C., 73 Scharenberg, A. M., 173
Rojas, J. R., 108, 109(14–16), 111(14–16), Schatz, M., 41
112(14–16), 114(15; 16), 117(16), 194 Scheer, U., 215
Roman, G., 428 Scheidtmann, K. H., 232
Rosa, E., 200 Scherthan, H., 237, 238(20), 240(20), 241(20),
Rose, A. B., 349 242(20), 246(20), 247(20)
Rose, S., 121, 219, 223, 232(16) Schiltz, L., 197
Roth, F. P., 317 Schiltz, R. L., 117(30), 119, 196, 223
Roth, S. Y., 106, 117(29), 119, 120, 131, 171, Schimmel, P. R., 140, 142(30)
189, 199, 223, 267 Schleucher, J., 128
Rougeulle, C., 247, 260, 406, 408(15), Schlichter, A., 120
420(15), 421 Schlisio, S., 316
Rowley, J. D., 295, 353 Schmatz, D. M., 174, 175
Royer, C. A., 4 Schmid, M., 255, 287, 387
Ruben, D. J., 166(33), 167 Schmidt, C. F., 62
Rubin, M. A., 270 Schmidt, E., 200
Rubin, R. L., 217 Schmidt, M. T., 173
Rudkin, G. T., 372, 374(2) Schnapp, B. J., 62
Ruiz, P., 200 Schneider, R., 237, 255, 270, 271(7), 283,
Rullmann, J. A., 125 283(7), 314, 349, 357(1), 360(1)
Rundlett, S. E., 289 Schnitzer, M. J., 62, 71(3)
Rupert, J. L., 420 Schnitzler, G., 64
Ruprecht, J., 38(9), 39 Schnoelzer, M., 197
Russell, D. W., 377, 384(15) Schofer, C., 237, 238(20), 240(20), 241(20),
242(20), 246(20), 247(20)
Schotta, G., 373, 387, 387(10),
S
389(10), 392(10)
Sachs, L. M., 197 Schramm, V. L., 173
Salvador, M. L., 174 Schreiber, J., 290, 317, 322(13), 333(13), 350,
Sambrook, J., 377, 384(15) 353(11), 357(11), 360
Samori, B., 76 Schreiber, S. L., 106, 171, 172(8), 174, 188,
Sanda, M. G., 270 189, 189(2), 199, 237, 270, 271(7), 283(7),
Santoso, B., 196 290, 301, 303, 314, 349, 354,
Santos-Rosa, H., 237, 270, 271(7), 283(7) 357(1), 360(1)
Sargent, D. F., 58, 59(9) Schreiber, S. S., 323
Sarma, K., 235, 248, 248(11), 254(11; 27), 255 Schroeder, M., 315
Sartorelli, V., 197 Schubert, U., 130
Sasaki, M., 405, 408(4), 417(4), 418(4) Schultz, P., 60
Sassone-Corsi, P., 171, 221, 223, Schurter, B. T., 272
232(15), 233(5) Schwartz, B. E., 372, 373(4), 374(4), 393,
Sathasivam, K., 200 393(4), 395
Sattler, M., 123, 128 Schweizer, D., 241, 242(24), 247(24), 260, 405
Saudek, V., 166(34; 35), 167 Schwienhorst, A., 176
Sauer, S., 237, 238(20), 240(20), 241(20), Scolnick, D. M., 128
242(20), 246(20), 247(20) Scott, M. P., 120
Sauve, A. A., 173 Sedat, J. W., 45, 47
Sawyer, W. H., 140 Seelig, H. P., 175
Saxton, W. O., 48 Segel, I. H., 187
444 author index
Seigneurin-Berny, D., 171 Singh, P. B., 149, 151(6), 168(6), 169(6), 237,
Seitz, V., 252, 262 242(22), 246(22), 248(22), 254(22), 387
Selenko, P., 387 Singh, S. B., 174
Selleck, W., 106(11), 107 Sistare, F. D., 28
Sendra, R., 174 Skelton, N. J., 165(31), 166
Sengupta, A., 252, 262 Sklenar, V., 166(34; 35), 167
Senn, H., 124 Slade, D., 290, 304
Serwer, P., 17, 18, 18(5; 6), 19(5; 6), 20(5; 6), Slama, J. T., 173, 177(23), 183(23), 185(23)
26(6), 27(5; 6; 11) Smith, A. G., 409
Serycheva, I. I., 41 Smith, B. O., 149, 151(7), 167(7)
Sewack, G. F., 272 Smith, C. L., 39, 41(11), 62, 65(8), 67(8),
Sewalt, R. G., 270 68(8), 71(8)
Shabanowitz, J., 231 Smith, D. L., 200
Shaffer, C. D., 373, 387, 387(8) Smith, J. S., 172(14), 173
Shah, V., 272, 282(16) Smith, P., 97, 103(32)
Shankaranarayana, G. D., 428 Smith, S. B., 62, 82, 92, 420
Sheardown, S. A., 420 Snyder, M., 290, 295, 317, 350, 353(10),
Sheppard, R. C., 273 354(10), 357, 359(16)
Sherriff, J., 237, 270, 271(7), 283(7) Sobel, R. E., 117(29), 119, 222, 223
Shi, H., 328, 333, 333(30) Soccio, R. E., 114, 189, 191(7), 193(7)
Shi, W., 237, 242(22), 246(22), 248(22), 254(22) Solomon, M. J., 336, 349
Shi, Y., 196, 197(25) Sommerville, J., 215
Shiloh, A., 173 Sopta, M., 394
Shinkai, Y., 237, 238(19), 245(19), 246(19), Sörensen, O. W., 124
249(19), 254(19) Sörensen, P., 124
Shoemaker, D. D., 290 Sorsa, V., 402
Shokat, K. M., 357 Sousa, K., 120
Shopland, L. S., 397, 403(13) Spector, D. L., 237, 238(21), 247, 247(21),
Shroff, R., 314 260, 405, 406, 408(15), 411, 413(23),
Shrogen-Knaak, M. A., 263 419(15), 421
Shundrovsky, A., 62, 67(5), 68(5) Speicher, D. W., 112, 114(23)
Sibilia, M., 237, 238(20), 240(20), 241(20), Spellman, P. T., 360
242(20), 246(20), 247(20) Spencer, V. A., 346
Siggia, E. D., 69(18), 72, 95 Sperier, A., 373
Silva, J., 248, 249(30), 254(30), 260, 405, Sperier, P., 373
406(9), 408(9), 421, 422(18) Spofford, J. B., 386
Silver, L. M., 372, 374, 376(1), 379, 380, Staes, M., 315
380(1), 394 Stahl, S. J., 136
Silver, P. A., 317 Stallcup, M. R., 231, 272
Silverman, N., 120 Stanik, V. H., 20
Simmler, M. C., 420 Starai, V. J., 172(14), 173, 174
Simon, I., 290, 314, 317, 322(13), 333(13), 350, Stebbins, J., 172(15), 173, 175(15), 177(15),
353(11), 357(11) 182(15), 183(15)
Simon, J. A., 248, 252, 254(28), 262, 290, 394 Steffan, J. S., 200
Simons, S. S., Jr., 28 Stelzer, G., 316
Simonson, T., 139 Stepaniants, S. B., 290
Simpson, M., 335, 348(5) Stephens, G. E., 372, 373, 387, 387(8), 403
Sinclair, D. A., 185, 386 Sternberger, L. A., 380
Singer, R. H., 411 Sterner, D. E., 106, 108, 109(15), 111(15),
Singh, N., 144 112(15), 114(15), 120
author index 445
Sternglanz, R., 112, 172(15; 16), 173, 175, Tafrov, S. T., 112, 172(15), 173, 175(15),
175(15), 177(15; 16; 23), 180(40), 177(15), 182(15), 183(15)
182(15), 183(15; 23), 185(23) Tagne, J. B., 314, 317
Steven, A. C., 42 Tahourdin, C. S., 214
Steward, R., 235, 248, 248(11), Takada, S., 395
254(11; 27), 255 Takagi, N., 405, 408(4), 417, 417(4),
Stewart, A. F., 119, 147, 152(4) 418(4), 422
Stoffler, G., 50 Takahashi, Y., 306, 318, 360
Stoffler-Meilicke, M., 50 Tamkun, J. W., 3, 120
Stollar, B. D., 212 Tan, E. M., 217
Stott, F. J., 149, 151(6), 168(6), 169(6) Tan, S., 65, 106(11; 12), 107, 120
Stott, K. M., 149, 167(5) Tang, H., 140, 255
Stoughton, R., 290, 313, 359 Tanner, K. G., 114, 172(16), 173, 177(16),
Stowell, M., 60 178, 194, 223, 232(15)
Strahl, B. D., 106, 121, 130(23), 131, 151, Tanny, J. C., 173, 182(19), 183(19)
199, 222, 231, 234(2), 235, 255, 269, 287, Tao, Y., 43
387, 388 Tastan, S., 420
Strahl-Bolsinger, S., 290, 292, 300(14), Tattersall, P., 316
304(14) Taunton, J., 174, 189, 199
Strausberg, R. L., 189, 317 Tavazoie, S., 290, 294(9), 303, 360
Strauss, W., 420 Taverna, S. D., 131, 132(8), 134(8), 140,
Strick, T. R., 76, 99, 101(6), 102(5) 140(8), 142(8; 29), 144(8)
Struhl, K., 235, 254, 289, 317, 336 Taylor, I. A., 270
Strutt, H., 349 Taylor, J. P., 199
Su, F., 173 Tempst, P., 120, 231, 235, 248, 252,
Subbaramaiah, K., 195 254(27; 28), 255, 262, 317, 332, 395, 408
Sugawara, O., 405, 408(4), 417(4), 418(4) Teng, G., 428
Suka, N., 209, 217(7), 289, 290, 293(1), 300, Teplyakov, A., 139
300(10), 301, 302(10), 303, 314(8), 315, Terragni, J., 306, 318, 360
349, 360(2) Thåström, A., 5, 12
Suka, Y., 209, 217(7), 289, 290, 293(1), Thijs, G., 315
300(10), 301, 302(10), 303, 314(8), 315, Thion, L., 316
349, 360(2) Thireos, G., 120
Sun, Z. W., 255, 287, 387 Thiru, A., 149, 151(7), 167(5; 7)
Surani, M. A., 235 Thiry, M., 33(26), 34
Sutton, A., 172(15), 173, 175(15), 177(15), Thoma, F., 79
182(15), 183(15) Thomas, J. O., 131, 151, 157(13), 170(13),
Sutton, C. A., 394 386, 387(28)
Svedberg, T., 161, 162(22) Thompson, B. A., 373, 387, 387(8)
Svoboda, K., 62 Thompson, C. M., 314, 317
Swat, W., 335 Thompson, L. M., 200
Switz, N. A., 82 Thompson, P. R., 115, 188, 191, 192, 193(13)
Syntichaki, P., 120 Thorne, A. W., 222, 336
Szyperski, T., 124 Thornton, J. M., 125, 210, 218(8)
Thuman-Commike, P. A., 38(10), 39
Tinoco, I., 62
T Tischendorf, G. W., 50
Tachibana, M., 237, 238(19), 245(19), Tjian, R., 127, 137, 144(20), 395
245(19), 249(19), 254(19) Tomschik, M., 76, 97, 103(32)
Taddei, A., 240, 271, 428 Tong, J. K., 290, 301, 303, 360
446 author index
Zeng, L., 121, 125(20), 126, 127(20), 128, Zhou, J., 106, 108, 109(15), 111(15; 16), 112,
128(20), 129(53), 130(20; 53), 137, 112(15; 16), 114(15; 16), 117(16), 120,
233, 304 189, 194, 199
Zhang, C. L., 198 Zhou, J. X., 108, 117(17)
Zhang, H. B., 128 Zhou, M.-M., 119, 121, 123, 125(20), 126,
Zhang, K., 255 127(20), 128, 128(20), 129(53), 130,
Zhang, M. Q., 315, 316, 328 130(20; 53), 137, 233, 270, 304
Zhang, T., 316 Zhou, S., 152, 395
Zhang, W., 43, 267 Zhou, X., 200
Zhang, Y., 106, 131, 132(8), 134(8), 140(8), Zhu, G., 123, 136
142(8), 144(8), 175, 221, 231, 233(4), 235, Zhu, Y. Z., 200
248, 252, 254(28; 29), 255, 260, 262, 272, Zink, D., 152
332, 405, 406(10), 310(10), 421, 422(19) Zlatanova, J., 73, 74, 76, 77(3), 78(3), 94,
Zhatnova, J., 62 95(29), 96(29), 97, 103(32)
Zheng, W., 190 Zvetkova, I., 248, 249(30), 254(30), 260, 405,
Zheng, Y., 188 406(9), 408(9), 421, 422(18)
Subject Index
A image acquisition, 79
materials, 78
Acetylation microarray surface preparation and sample
applications, 302–304 deposition, 79
chromatin immunoprecipitation instrumentation, 76–78
cell growth and harvesting, 292 principles, 73–74
cross-linking AUT gel electrophoresis, see
double cross-linking, 294–295 Acid-urea-triton gel electrophoresis
formaldehyde, 292
immunoprecipitation reaction, 293
polymerase chain reaction, 295–297 B
sonication, 292–293
troubleshooting, 293–294 Brahma, bromodomain structure, see
DNA microarray analysis Nuclear magnetic resonance
data quantification, normalization, and Bromodomain structure, see Nuclear
analysis, 299–302 magnetic resonance
Klenow labeling of probe and
hybridization, 297–299
C
yeast intergenic microarray
preparation, 302 ChIP, see Chromatin immunoprecipitation
principles, 290–292 Chromatin, see also Histone; Nucleosome
Acid-urea-triton gel electrophoresis, histone agarose multigel electrophoresis analysis
deacetylase assay for inhibitor of structure, see Agarose multigel
characterization, 201–202 electrophoresis
AFM, see Atomic force microscopy antibodies, see also Histone antibodies
Agarose multigel electrophoresis applications, overview, 209–210
chromatin structure analysis histone antigen interactions, 209–212
composition analysis of genomic immunogen preparation
fragments, 27–28 dehistonized chromatin, 214–215
data interpretation, 28–29 high mobility group proteins, 215–216
flexibility assay, 27 histone fractions, 213–214
overview, 17–18, 20–21 nucleic acids, 216
secondary chromatin structure nucleosomes, 214
formation, 26 synthetic histone peptides, 212–213
data analysis, 25–26 preparation, 216–217
electrophoresis, 24 specificity and affinity, 217–220
equipment, 21 electron microscopy three-dimensional
gel preparation, 22, 24 structure determination
principles, 18–19 adhesion to support films, 31
Southern blot, 25 buffer components, 30
AME, see Agarose multigel electrophoresis concentration of sample, 30
Atomic force microscopy cryomicroscopy, 36
chromatin fiber imaging electron tomography
glutaraldehyde fixation, 78–79 alignment and reconstruction, 46–47
449
450 subject index
methyl-histone H3 antibody R
chromatin immunoprecipitation
analysis, 286 RSC chromatin remodeling complex,
random-primer polymerase chain reaction electron microscopy
for genome-wide analysis of histone nucleosome interaction imaging, 55–59
modifications in yeast with chromatin sample requirements and preparation,
immunoprecipitation and DNA 48–50
microarrays, 354–356 specimen evaluation and data collection,
real-time quantitative polymerase chain 50–51
reaction in chromatin three-dimensional structure reconstruction
immunoprecipitation analysis initial calculations, 53
amplification reactions, 344–345 interpretation, 53–55
calculations, 345 two-dimensional image analysis and
fluorescent probes, 344 classification, 51–52
primer design, 342–343 unstained sample studies, 59–62
principles, 343–344
Polytene chromosome
amplification in development, 372, 393 S
immunofluorescence microscopy of Sir2
Drosophila salivary gland activity assays
chromosome proteins
charcoal-binding assay, 180–182
advantages and limitations, 402–404 high-performance liquid
antibody specificity testing, 395 chromatography of peptides,
antibody staining, 379–381, 400–401 178–180
custom P-element studies, 390, 392
kinetic parameterr analysis, 186–187
fixation, 374, 376–377 nicotinamide assays
fly culture, 399 catalyzed exchange assay, 183–195
genetic manipulation of flies, 373, 385, inhibition assays, 185
396–397
substrate preparation, 177
heat shocking of larvae, 399 thin-layer chromatography assay,
image analysis, 401–402 182–183
microscopy, 381–382, 401
catalytic mechanism, 173
multiple staining of heterochromatin functions, 172–173
proteins and histone modifications, Southern blot, agarose multigel
386–388 electrophoresis, 25
mutant fly studies, 389–390
non-melanogaster species analysis,
388–389
T
overview, 372–374
prospects, 392–393 TEM, see Transmission electron microscopy
protein colocalization studies, Transmission electron microscopy
394–395 chromatin three-dimensional structure
salivary gland dissection, 399 determination
sensitivity enhancement, 396–397 adhesion to support films, 31
spatial resolution, 395–396, 402 buffer components, 30
squashing, 377–379, 400 concentration of sample, 30
troubleshooting, 382–385 cryomicroscopy, 36
structure, 393–394 electron tomography
Protein arginine methyltransferases, see alignment and reconstruction, 46–47
Histone methyltransferase data collection, 45–46
458 subject index