Histone Acetylation and Deacetylation - CiteSeerX

4 downloads 0 Views 569KB Size Report
The acetylation isoforms of histone H4 from butyrate- treated HeLa cells were separated by C4 reverse-phase high pressure liquid chromatography and by ...
Research

Histone Acetylation and Deacetylation IDENTIFICATION OF ACETYLATION AND METHYLATION SITES OF HeLa HISTONE H4 BY MASS SPECTROMETRY*

Kangling Zhang‡§¶, Katherine E. Williams§, Lan Huang§, Peter Yau储, Joseph S. Siino储, E. Morton Bradbury储**, Patrick R. Jones‡, Michael J. Minch†‡, and Alma L. Burlingame§‡‡ The acetylation isoforms of histone H4 from butyratetreated HeLa cells were separated by C4 reverse-phase high pressure liquid chromatography and by polyacrylamide gel electrophoresis. Histone H4 bands were excised and digested in-gel with the endoprotease trypsin. Matrix-assisted laser desorption ionization time-of-flight mass spectrometry was used to characterize the level of acetylation, and nanoelectrospray tandem mass spectrometric analysis of the acetylated peptides was used to determine the exact sites of acetylation. Although there are 15 acetylation sites possible, only four acetylated peptide sequences were actually observed. The tetra-acetylated form is modified at lysines 5, 8, 12, and 16, the tri-acetylated form is modified at lysines 8, 12, and 16, and the di-acetylated form is modified at lysines 12 and 16. The only significant amount of the mono-acetylated form was found at position 16. These results are consistent with the hypothesis of a “zip” model whereby acetylation of histone H4 proceeds in the direction of from Lys-16 to Lys-5, and deacetylation proceeds in the reverse direction. Histone acetylation and deacetylation are coordinated processes leading to a non-random distribution of isoforms. Our results also revealed that lysine 20 is dimethylated in all modified isoforms, as well as the nonacetylated isoform of H4. Molecular & Cellular Proteomics 1:500 –508, 2002.

The basic structural unit of eukaryotic chromosomes is a DNA䡠protein complex called the nucleosome. The nucleosome consists of a DNA molecule associated with a histone octamer comprised of pairs of the core histones H2A, H2B, H3 and H4. The nucleosomes are joined by linker DNA and histone H1 to form chromatin. Each core histone has a globular region, the histone fold domain, which is involved in histone-histone interactions (1), and the wrapping of DNA From the ‡Department of Chemistry, University of the Pacific, Stockton, California 95211, §Department of Pharmaceutical Chemistry, Mass Spectrometry Facility, University of California, San Francisco, California 94143, 储Department of Biological Chemistry, University of California, Davis, California 95616, and **MS M-888 B-2, B-Division, Los Alamos National Laboratory, Los Alamos, New Mexico 87545 Received, May 24, 2002, and in revised form, July 9, 2002 Published, MCP Papers in Press, July 16, 2002, DOI 10.1074/ mcp.M200031-MCP200

500

Molecular & Cellular Proteomics 1.7

around the nucleosome core (2). The N-terminal “tail regions” extend outside of the nucleosome particle where they can interact with DNA and with other regulatory proteins or transcription factors (2– 4). These tails are essential but appear partially redundant. At least one of the two tails for both the H2A-H2B and H3-H4 pairs must be intact to maintain viability of yeast cells (5). The tails account for 28% of the core histone sequences and are extremely basic because of a high proportion of lysine and arginine. In the 2.8-Å nucleosome crystal structure, the electron densities for these tails are largely not observed (6). Presumably their binding sites lie outside of the core particle. Several studies indicate that these tails do not contribute significantly to the primary wrapping of DNA in the nucleosome. Trypsinized nucleosome core particles are just as stable as intact particles with respect to perturbations in temperature, high salt concentrations, and accessibility to DNase I (7, 8). Neutron scattering studies have shown that hyperacetylation of histones has little or no effect on core particle structure in solution (9). However, the histone tail sequences are highly conserved, and the reversible acetylation of the ⑀-amino groups of specific lysine residues has been implicated in key regulatory events (3, 10 –15). The acetylation of histone H4 is restricted to lysine 5, 8, 12, and 16 (16). Another known post-translational modification is methylation of lysine 20 (17), which precludes acetylation at this site (18). Random histone acetylation would yield four mono-acetylated isoforms, six di-acetylated isoforms (Lys-5/ Lys-8, Lys-5/Lys-12, Lys-5/Lys-16, Lys-8/Lys-12, Lys-8/Lys16, and Lys-12/Lys-16), four tri-acetylated isoforms (Lys-5/ Lys-8/Lys-12, Lys-5/Lys-12/Lys-16, Lys-8/Lys-12/Lys-16, and Lys-5/Lys-8/Lys-16), and one tetra-acetylated isoform. Table I shows the N-terminal sequence of H4, which contains all four acetylation sites (Lys-5, -8, -12, and -16) and the single methylation site at lysine 20. There are clearly functional differences among lysine residues. However, the specific role of each of these forms will remain unclear until measurable differences in the chromatographic, electrophoretic, and mass spectrometric properties of these forms can be correlated with cellular events (19). This study is directed toward that goal. Histone acetylation is a very specific phenomenon with various isoforms playing distinct roles (13). Acetylation is a dynamic phenomenon with the steady state mediated by the

© 2002 by The American Society for Biochemistry and Molecular Biology, Inc. This paper is available on line at http://www.mcponline.org

Research

FIG. 2. MALDI-TOF spectrum of histone H4 digested by trypsin. The spectrum was calibrated by internal two-point calibration, using bradykinin (20 fmol; peak 1060.569) and trypsin autolysis product (peak 2163.057). Peaks were assigned to peptide sequences of histone H4 by matching the measured masses with the calculated masses (cf. Table I).

FIG. 1. SDS-PAGE of HPLC fraction containing HeLa histone H4 and H2A.

opposing activities of histone acetyltransferases (HATs)1 and deacetylases. These activities involve large regulatory complexes that are capable of responding to specific DNA sequences and can contain transcription factors, regulatory ligands, and signal transduction and cell cycle proteins. Acetylation plays a role in nucleosome assembly. Newly synthesized histone H4 is di-acetylated in the cytoplasm at Lys-5 and Lys-12 by B-type histone acetyltransferase (HAT B) (20). CAF-1 (chromatin assembly factor 1) is a complex of proteins that deposits Lys-5, Lys-12 di-acetylated H4 into chromatin. It has also been found that human CAF-1 readily assembles newly synthesized H3 and H4 onto replicating DNA in vitro. Presumably the specific acetylation patterns of nascent histones are required for their assembly in vitro (21). The acetylation pattern found in mature nucleosomes differs from that of newly incorporated histones. Increased acetylation is generally correlated with transcriptionally active or poised genes. Histone acetylation within nuclei is effected by A-type histone acetyltransferases (HAT A) that are likely to participate in gene activation. An in-gel HAT assay was used to purify and clone the cDNA for the major macronuclear HAT A, p55, from Tetrahymena (22). The gene encoding this enzyme is homologous to the yeast transcriptional co-activator gene Gcn5. The yeast protein Gcn5p, with ADA2 and ADA3, interacts with enhancer binding factors thereby establishing a direct mechanistic relationship between acetylation and gene activation (23). The yeast protein has also been shown to possess HAT activity. Gcn5p and p55 both contain a bromodomain (12), which has been shown to 1

The abbreviations used are: HAT, histone acetyltransferase; HPLC, high pressure liquid chromatography; MALDI, matrix-assisted laser desorption ionization; PSD, post-source decay; ESI, electrospray; MS/MS, tandem mass spectrometry; CID, collision-induced dissociation.

be important for the assembly and activity of multisubunit transcriptional activation complexes (12). The bromodomain may tether HAT A to specific chromosomal sites, linking histone acetylation and gene activation (12, 22, 24). The Gcn5p䡠ADA complex interacts functionally with the SWI䡠SNF complex, which is part of the RNA polymerase II holoenzyme (12, 25). Further evidence for the connection between acetylation and transcription activation is the finding that H4 shows increased Lys-16 acetylation in the hyperactive X chromosome of male Drosophila (26). In the inactive X chromosome of female flies, H4 is hypoacetylated (27). Immunochemistry experiments demonstrate that H4 in transcriptionally inactive heterochromatin is hypoacetylated predominately in mammals. H4 histones from the heterochromatins of Drosophila (28) and yeast (29) are acetylated only at Lys-12, suggesting that this modification is important in gene silencing in these species. Histone acetylation and deacetylation may play critical roles in cell cycle procession (3). The natural synchrony of Physarum polycephalum, which has no G1 phase, enabled Matthews (30) to distinguish between acetylation specific for S phase and G2 phase. H4 in growing cells was found to be acetylated at Lys-5, Lys-8, Lys-12, and Lys-16. In G2 phase, however, the turnover was highest at Lys-8 whereas in S phase turnover was highest at Lys-5. As the above examples illustrate much will be gained by having methodology that is easily able to identify the relative populations of acetylation isoforms as a function of cellular events and gene location. Allis and co-workers (31, 32) used [3H]lysine pulse labeling and Edman microsequencing to determine the acetylation sites of newly synthesized H4 in many organisms including Tetrahymena, Drosophila, and humans. Other microsequencing methods have been used by Couppez et al. (33), Matthews (30), Thorne et al. (34), and Waterborg (35). Couppez et al. (33) observed that the principle tri-acetylated form of cuttlefish H4 is Lys-5/Lys-12/Lys-16 and that the principle di-acetylated and mono-acetylated forms are Lys-5/ Lys-12 and Lys-12 suggesting that Lys-5 is acetylated after

Molecular & Cellular Proteomics 1.7

501

Research

TABLE I Peptide fragments of histone H4 observed by MALDI-TOF mass spectrometry Kac, acetylated lysine residue; Kme2, dimethylated lysine residue. AA residues

Measured (M⫹H)⫹

Calculated (M⫹H)⫹

⌬M (Da)

13–17

530.26

530.26

0.01

20–23

543.36

543.36

0.00

9–17

927.53

927.53

0.00

60–67 46–55 6–17

989.57 1180.60 1211.70

989.57 1180.61 1211.68

0.00 0.01 0.02

68–78 24–35 45–55 4–17

1290.60 1325.80 1336.70 1396.80

1290.64 1325.74 1336.71 1396.80

0.04 0.06 0.01 0.00

4–17

1438.80

1438.80

0.00

24–36

1481.80

1481.84

0.04

Lys-12 (33). In calf thymus H4 (36), Lys-16 is the only monoacetylated position, and all di-acetylated forms also involve Lys-16, implying that Lys-16 is acetylated first. Turner used specific antibodies for acetylated forms of H4 to conclude that Lys-5 and Lys-12 positions are under-used in mono-acetylated H4 from a number of mammalian cell lines and that Lys-8 and/or Lys-16 are the first position(s) to be acetylated (36). Treatment of actively growing cells with histone deacetylase inhibitors such as sodium butyrate shifts the equilibrium in the direction of acetylation of histones (37). Thorne et al. (34) have reported that for H4 from pig thymus and from sodium butyrate-treated HeLa cells Lys-16 is the predominant mono-acetylation site, but further addition of acetyl groups is less specific and progresses though sites Lys-12, Lys-8, and Lys-5 in an N-terminal direction. For example from butyratetreated HeLa cells, the Lys-8/Lys-12/Lys-16 isoform was found to be more than twice as abundant as the Lys-5/Lys-12/Lys-16 and Lys-5/Lys-8/Lys-16 isoforms. In all the above cases the data were complicated by the kinetics of acetylation-deacetylation reactions and the procedures for histone extraction and separation, which often involved multiple steps of column chromatography. HPLC and/or electrophoresis followed by mass spectrometric characterization of the separated components shows great promise for revealing such patterns. Earlier, Edmonds et al. (38) reported the use of electrospray ionization mass spectrometry and tandem mass spectrometry to evaluate the sequence and modifications of histone H3. We report here the use of mass spectrometric methods to identify the acetylation sites of histone H4 to clarify the complicated and somewhat contradictory picture regarding the relative abundances of H4 acety-

502

Molecular & Cellular Proteomics 1.7

Sequence

16 GGAKacR 20 Kme2 VLR 12 16 GLGKacGGAKacR VFLENVIR ISGLIYEETR 8 12 16 GGKacGLGKacGGAKACR DAVTYTEHAKR DNIQGITKPAIR RISGLIYEETR 8 12 16 GKGGKacGLGKacGGAKacR 5 8 12 16 GKacGGKacGLGKacGGAKacR DNIQGITKPAIRR

lation isoforms. Our mass spectrometric approach shows directly, effectively, and with high sensitivity that the acetylation of H4 in HeLa cell nuclei proceeds from lysine 16 to lysine 5. These results suggest the hypothesis of a “zip” model whereby acetylation of histone H4 proceeds in the direction from Lys-16 to Lys-5, and deacetylation proceeds in the reverse direction. Our results also revealed that lysine 20 is ⑀-amino-di-methylated in all the acetylated isoforms, as well as the non-acetylated isoform of H4. MATERIALS AND METHODS

Cell Culture and Histone Separation—HeLa S3 (originally from ATCC) were cultured in RPMI 1640 supplemented with 10% calf serum (Invitrogen) as described (39) or purchased from National Cell Culture Center. Hyperacetylated histones were obtained by treating the cultures with 7 mM sodium butyrate (Mallinckrodt) for 24 h prior to harvest. Cells were washed in phosphate-buffered saline containing 10 mM sodium butyrate, and histones were extracted from isolated nuclei with 0.4 N H2SO4 and separated using reverse-phase HPLC Vydac C4 column (Vydac C4 TP214; 10-␮m beads, 300-␮m pores) on a PerkinElmer Life Sciences Bio 410 system with a diode array detector. A flow rate of 10 ml/min with a gradient of water and CH3CN containing 0.1% trifluoroacetic acid was used. Fractions containing H4 (co-eluted with minor H2A) were confirmed by SDS-PAGE (see Fig. 1) and stained using Coomassie Blue (Novex). In-gel and In-solution Enzymatic Digestion—Histone H4 bands were excised from the gel and digested with trypsin (40). For endoproteinase Asp-N digestion, 5 ␮g of purified histones were dissolved in 20 ␮l of 25 mM NH4HCO3 and incubated at 37 °C overnight with 300 ng of Asp-N. After adding 5 ␮l of stop solution (50% acetonitrile, 45% water, 5% trifluoroacetic acid), digested peptides were concentrated using a Speedvac. Mass Spectrometric Analysis of Peptides—Monoisotopic mass values of all peptides were measured by MALDI using a Voyager DE-STR biospectrometry workstation (PerSeptive Biosystems, Inc.) with de-

Research

TABLE II Calculated Possible Trypsin Digest Peaks of N-terminal Histone H4 Isoforms Mono-acetylated

Di-acetylated

Tri-acetylated

Tetra-acetylated

MH⫹

Trypsin-digested Sequences

5 GKacGGK 8 GGKacGLGK 12 GLGKacGGAK 16 GGAKacR 5 8 GKacGGKacGLGK 5 12 GKacGGK, GLGKacGGAK 5 16 GKacGGK, GGAKacR 8 12 GGKacGLGKacGGAK 8 16 GGKacGLGK, GGAKacR 12 16 GLGKacGGAKacR 5 8 12 GKacGGKacGLGKacGGAK 5 8 16 GKacGGKacGLGK, GGAKacR 5 12 16 GKacGGK, GLGKacGGAKacR 8 12 16 GGKacGLGKacGGAKacR 5 8 12 16 GKacGGKacGLGKacGGAKacR

488.3 658.4 729.4 530.3 885.5 488.3, 729.4 (1198.7)a 488.3; 530.3 (999.6) 1013.6 658.4; 530.3 (1169.7) 927.5b 1240.7 885.5; 530.3 (1396.8) 488.3, 927.5 (1396.8) 1211.7b 1438.8b

a

Numbers in parenthesis are mass values of the peptide sequences uncleaved at indicated site (comma). The peptide sequences were established by nano-ESI/MS/MS (see Figs. 5–7). The highlighted mass values were observed by MALDI-TOF mass spectrometry (see Fig. 2 and Table I). b

FIG. 3. MALDI-PSD spectrum of parent ion at m/z 543.35 Da.

layed extraction operated in the reflectron mode. A resolution of 5000 to 10,000 full-width at half-maximum was achieved within the mass range of 1000 to 3000. Between 100 and 256 accumulations were analyzed for each spectrum, and peptides were observed as (M⫹H)⫹ in the positive mode. Post-source decay (PSD) spectra were recorded for selected ions. Peptides (1/10 of sample) were co-crystallized with equal volumes of a saturated ␣-cyano-4-hydroxycinnamic acid matrix (Hewlett Packard) in 50% acetonitrile, 0.1% trifluoroacetic acid. All MALDI spectra were calibrated internally using trypsin autolysis products and/or calibrated externally by an added peptide standard. Nano-ESI Tandem Mass Spectrometry Analysis of Peptides—Tan-

dem mass spectra of peptides were obtained on a quadrupole orthogonal acceleration time-of-flight mass spectrometer with a quadrupole collision cell (QSTAR; Sciex, Toronto, Ontario, Canada) with an external nanoelectrospray ion source (Protana A/S, Odense, Denmark). Mass resolution was obtained routinely in the range of 7000 to 12000 (for both conventional mass spectrometric and MS/MS modes of operation), and a mass measurement accuracy of at least 0.02 Da with external calibration was achieved. Approximately 2 ␮l of sample was loaded into a nanoelectrospray tip. Conventional mass spectra were first obtained to measure mass values (parent ions) of trypsindigested peptides and to assign their charge states from observation of stable isotope spacing. Then parent ions of interest were selected for sequence analysis by tandem mass spectrometry. RESULTS

Identification of Acetylated Histone H4 by MALDI-TOF Mass Spectrometry—Histone H4 samples were resolved by SDSPAGE and visualized using Coomassie Blue stain. The single histone H4 band containing all acetylated isoforms was excised and digested with trypsin. The resulting tryptic peptides were analyzed by MALDI-TOF mass spectrometry to determine their individual mass values (Fig. 2). The specific peptide sequences of histone H4 were assigned to the mass values in

Molecular & Cellular Proteomics 1.7

503

Research

FIG. 4. Nano-ESI-MS/MS spectrum of precursor ion at 719.94 Da. Interpretation of this CID spectrum established the tetra-acetylmodified sequence covering residues 4 –17 as GKacGGKacGLGKacGGAKacR. The ⑀-amino function of residues 16, 12, 8, and 5 are acetylated. In the figure, ions labeled as a, b, c, and their corresponding water or ammonia loss ions are N-terminal sequence ions whereas those labeled as y and their water loss ions are C-terminal sequence ions (fragmentation directions are shown above the spectrum). Ions labeled as i are the internal fragmentation ions: the ion at m/z 356.20 corresponds to the fragment KacGGA; 427.31 corresponds to the fragment GGKcGL-28, KacGGKac, GKacGLG-28, GLGKacG-28, or LGKacGG-28; and 455.29 corresponds to the fragment KacGGKac, GGKacGL, GKacGLG, LGKacGG, or GLGKacG. The signal at m/z 126.10 is the immonium ion of ⑀-aminoacetyl lysine.

Fig. 2 by matching the measured masses with expected calculated mass values. The resulting assignments are listed in Table I, in which the acetylation sites contained in each specific peptide sequence are highlighted. Whereas unacetylated lysine residues are susceptible to cleavage by trypsin, acetylated lysine residues are resistant. The calculated mass values for the peptide sequences corresponding to the various acetylated isoforms are listed in Table II. By comparing the measured mass values of the peptide sequences that contain putative acetylation sites (see Fig. 2 and Table I) with the mass values of all possible trypsin-digest peaks of N-terminal portion of histone H4 (see Table II), we find that only four of the 15 possible acetylated peptide sequences were observed in the MALDI-TOF spectrum of HeLa histone H4. The mass values of these four peaks were 1438.80, 1211.70, 927.53, and 530.26 Da. These masses correspond to the tetra-acetylated peptide sequence GK5acGGK8acGLGK12acGGAK16acR, the tri-acetylated peptide sequence GK8acGLGK12acGGAK16acR, the di-acetylated peptide sequence K12acGGAK16acR, and the mono-acetylated peptide sequence GGAK16acR. The peak at mass value 1396.80 Da corresponds to peptide sequence

504

Molecular & Cellular Proteomics 1.7

G5KGGK8acGLGK12acGGAK16acR from the tri-acetylated isoform of histone H4 in which the unmodified lysine residue 5 was not cleaved during the trypsin digestion. Upon prolonged tryptic digestion of this component, the peak at 1396.80 eventually disappeared whereas the relative intensity of 1211.70 increased proportionately. The peak at mass value 543.35 Da is assigned to peptide sequence 20 –23, KMe2VLR, where lysine 20 is modified by two methyl groups. The sequence was established by MALDI-PSD (see Fig. 3). Acetylation Sites Established Unambiguously by Nano-ESI/ MS/MS—The covalently modified peptides were subjected to collision-induced dissociation (CID) using Nano-ESI/MS/MS to establish the acetylation sites. The peptide sequence corresponding to the precursor ion at mass value 719.94 Da was established by nano-ESI/MS/MS to be the 5, 8, 12, 16-tetraacetyl analog of peptide sequence 4 –17 of histone H4: 4 GKacGGKacGLGKacGGAKacR. As shown in Fig. 4, the most abundant CID fragment ions (from nano-ESI/MS/MS) are b, y, and y-NH3 and b-H2O ions and a series of internal ions (labeled as i in the figure). A few a-type ions and one c-type ion at 359.2 Da were also observed. The unique immonium ion of 126 Da (major) and 143.1 Da (minor) are assigned unam-

Research

FIG. 5. Nano-ESI-MS/MS spectrum of precursor ion at 606.37 Da. Interpretation of this CID spectrum established the tri-acetyl-modified peptide sequence covering residues 6 –17 as GGKacGLGKacGGAKacR. In the figure, ions labeled as i are internal fragmenation ions: the ion at m/z 228.15 corresponds to the fragment GKac, KacG, or GLG; and 313.01 corresponds to the fragment KacGL-28 or LGKac-28. Signals at m/z 126.10 and 143.14 are immonium ions derived from ⑀-acetyl lysine residue.

biguously to acetylated lysine (Kac). This CID spectrum excludes the possible alternative to peptide fragment 79 –91, KTVTAMDVVYALK, which is isobaric with the tetra-acetylated peptide sequence 4 –17. Using a similar nano-ESI tandem mass spectrometry strategy, the nature of precursor ion mass at 606.37 Da was established as peptide sequence 6 –17, 6GGKacGLGKacGGAKacR, where lysine residues 8, 12, and 16 are all acetylated (Fig. 5). The other possible assignment of molecular mass 1211.71 Da to the uncleaved peptide sequence 92–102, RQGRTLYGFGG (mono-isotopic mass is 1211.62), is excluded by this CID spectrum. The lack of observation of this particular peptide could be because of digestion at Arg-95 to produce two smaller peptides 92–95 and 96 –102. The precursor mass at 464.36 Da can be assigned either as peptide sequence 4 –12, 4GKacGGKacGLGK (calculated mono-isotopic mass 927.52), where lysine 5 and 8 are acetylated, or peptide sequence 9 –17, 9GLGKacGGAKacR (calculated mono-isotopic mass 927.53), where lysine 12 and 16 are acetylated. The results from the nano-ESI tandem mass spectrometry experiment (shown in Fig. 6) revealed that this component matches unambiguously the peptide sequence 9 –17 rather than that of the sequence 4 –12. Although the mass of 927.54 Da indicates that lysine 12 and 16 are acetylated, the

origin of this peptide could not be confirmed. It could arise from either the di-acetylated isoform of histone H4 (lysine 12 and 16 are acetylated) or the tri-acetylated isoform of histone H4 in which lysine 5, 12, and 16 are acetylated, and lysine 8 is not acetylated and cleaved by trypsin (see Table II). However, the peptide sequence containing acetylated Lys-5 with a mono-isotopic mass of 488.3 Da corresponding to peptide sequence 4 – 8 (4GKacGGK) was not observed (Fig. 2) so that the tri-acetylated isoform of histone H4 at sites Lys-5, Lys-12, and Lys-16 is less likely to be the correct assignment. Histone H4 Is Di-methylated at Lysine Residue 20 in All Isoforms—To confirm that lysine 20 is modified by two methyl groups in all forms of histone H4, histone H4 was digested with Asp-N, which liberated peptide 1–23. As shown in Fig. 7, masses of 2430.4, 2472.4, 2514.5, 2557.2, and 2598.5 correspond respectively to un-, mono-, di-, tri-, and tetra-acetylated peptide isoforms (42 Da is the incremental mass of an acetyl group). Based on the fact that the measured mass values are 28 mass units higher than mass values of the unmethylated sequence, it is reasonable to conclude that lysine 20 must be modified by two methyl groups. Moreover, di-methylation of lysine residue 20 occurs in all H4 isoforms with different degrees of acetylation. That lysine residues 5, 8, 12, and 16 are acetylation sites has been proved by trypsin

Molecular & Cellular Proteomics 1.7

505

Research

FIG. 6. Nano-ESI-MS/MS spectrum of precursor ion at 464.36 Da. Interpretation of this CID spectrum established the di-acetyl-modified peptide sequence covering residues 9 –17 as GLGKacGGAKacR. Ions labeled as i are the internal fragmentation ions: the ion at m/z 285.18 corresponds to the fragment KacGG or GKacG; 342.26 corresponds to the fragment GKacGG; and 356.23 corresponds to the fragment KacGGA or GGAKac. Signals at m/z 126.10 and 143.14 represent immonium ions of acetyl lysine. Signals at m/z 84.09 and 86.11 are immonium ions corresponding to the lysine and arginine residues, respectively.

20 as the potential methylation site. The di-methylation at lysine residue 20 was confirmed by the identification of the peptide sequence 20 –23, KMe2VLR, sequenced by MALDIPSD (see Fig. 3). DISCUSSION

FIG. 7. Mass spectrum of histone H4 digested with Asp-N in solution. Mass values at m/z 2430.40, 2472.40, 2514.50, 2557.20, and 2598.50 Da correspond to un-, mono-, di, tri- and tetra-acetylated peptide sequence 1–23 (labeled as aa1–23Ac0 – 4) of histone H4 cut by enzyme Asp-N. Histone H4 is di-methylated at lysine 20 in all isoforms with different degrees of acetylation.

digestion and identified/confirmed by MALDI-TOF and nanoESI tandem mass spectrometry and are therefore excluded as possible sites for methylation. This leaves only lysine residue

506

Molecular & Cellular Proteomics 1.7

We have shown that tandem mass spectrometry can be used to identify the specific sites of modification of individual histone acetylation isoforms from a mixture and that the method is sufficiently sensitive to gain sequence information from samples extracted from electrophoresis gels. This is a remarkable improvement in sensitivity and ease of assignment over earlier microsequencing methods. It also overcomes the problems associated with N-terminal acetylation of H4 Ser-1, which blocks Edman degradation. These studies have established that there is a clear sequential pattern to the multiple acetylation of H4. For butyratetreated HeLa cells the only detectable acetylated forms are as follows: (i) the tetra-acetylated form is acetylated at lysine 5, 8, 12, and 16; (ii) the tri-acetylated form is acetylated at lysine 6, 12, and 16; (iii) the di-acetylated form is acetylated at lysine 12 and 16; and (iv) the only significant amount of mono-acetylated form is modified only at position 16. This provides ex-

Research

FIG. 8. A zip model describing the acetylation and deacetylation process of histone H4. HDAC, histone deacetylase.

perimental evidence for the observations of Couppez et al. (33), Turner et al. (28), and Thorne et al. (34), although these results are for total histone extracted from HeLa cells, treated with a deacetylase inhibitor, and thus describe the average acetylation pattern of histones. The levels and positions of acetylation of histones at individual genes may differ considerably and are governed by both HATs and by histone deacetylase. The actions of these two enzymes must be coordinated at most genes if a non-random distribution of isoforms for the entire cell is observed. The HPLC/gel electrophoresis/mass spectrometry approach may be applied to a wider range of cells and to specific chromosomes or chromatin domains and to cell cycle events to clarify such processes. As a working hypothesis we suggest a zip model (see Fig. 8) where histone acetylation starts at lysine 16 and continues until the four sites are acetylated whereas histone deacetylation must proceed in the reverse direction. Our mass spectrometric results (Fig. 7) revealed that lysine 20, in the N terminus of histone H4, is di-methylated and that methylation exists in all isoforms of histone H4 regardless of the degree of acetylation. It would be interesting to see whether this dimethylation site plays any role in acetylation. Does it serves as part of the initial HAT recognition and binding site? Is there something special about the pattern of acetylation (16 3 12 3 8 3 5)? The acetylation of histones results in the loss of the positive charges on lysine residues located in the core histone N termini, which would weaken the association between histone N-terminal domains and DNA. Thermal denaturation studies revealed that acetylation markedly reduces the binding constant of the H4 N terminus to DNA by six orders of magnitude in dilute buffer containing 5 mM Tris-HCl (41). Protection experiments with a model system consisting of phage ␭ DNA as a substrate, and H4 with different degrees of acetylation as the protective agent demonstrated that non-

acetylated H4 and mono-acetylated H4 cause similar protection from StuI, whereas di-, tri-, and tetra-acetylation of histone H4 result in cleavage of the DNA (42). Thus, nonacetylated H4 and mono-acetylated H4 have a comparable degree of protection of the DNA whereas di-acetylated, triacetylated, and tetra-acetylated H4 show significantly reduced protection against cleavage. Physical studies are required to determine whether the site of acetylation is as important as the number of positive charges neutralized. It is easy to speculate that the site of histone acetylation plays an important role in the binding of histones with transcription factors, because such specificity in protein-protein interaction is not uncommon. The non-acetylated histones are likely to adopt specific secondary structures when bound to DNA facilitating higher order chromatin condensation but precluding binding transcription factors or other specific chromatinassociated proteins whereas acetylation has been shown to favor transcription factor binding (43– 46). Acknowledgment—We acknowledge the National Cell Culture Centre for the growth of HeLa cells. * This work was supported in part by National Institutes of Health National Center for Research Resources Biomedical Research and Technology Program Grant RR01614 (to A. L. B.) and United States Department of Energy Grant DE-FG03-01ER 63070 (to E. M. B.). The costs of publication of this article were defrayed in part by the payment of page charges. This article must therefore be hereby marked “advertisement” in accordance with 18 U.S.C. Section 1734 solely to indicate this fact. This work is dedicated to the memory of Dr. Michael J. Minch. † Deceased. ¶ Present address: Dept. of Chemistry, University of Riverside, Riverside, CA. ‡‡ To whom correspondence should be addressed. E-mail: [email protected].

Molecular & Cellular Proteomics 1.7

507

Research

REFERENCES 1. Arents, G., Burlingame, R. W., Wang, B. C., and Love, W. E. (1991) The nucleosomal core histone octamer at 3.1 Å resolution: a tripartate protein assembly and a left-handed superhelix. Proc. Natl. Acad. Sci. U. S. A. 88, 10148 –10152 2. Luger, K., Ma¨ der, A. W., Richmond, R. K., Sargent, D. F., and Richmond, T. J. (1997) Crystal structure of the nucleosome core particle at 2.8 Å resolution. Nature (London) 389, 251–260 3. Bradbury, E. M. (1992) Reversible histone modifications and the chromosome cell cycle. Bioessays 14, 9 –16 4. Bane´ ras, J. L., Martin, A., and Parello, J. (1997) The N tails of histones H3 and H4 adopt a highly structured conformation in the nucleosome. J. Mol. Biol. 273, 503–508 5. Ling, X., Harkness, T. A., Schultz, M. C., Ficher-Adams, G., and Grunstein, M. (1996) Yeast histone H3 and H4 amino termini are important for nucleosome assembly in vivo and in vitro: redundant and position independent functions in assembly but not in gene expression. Genes Dev. 10, 686 – 699 6. Lugar, K., and Richmond, T. J. (1998) The histone tails of the nucleosome. Curr. Opin. Genet. Dev. 6, 140 –146 7. Ausio, J., Dong, F., and van Hold, K. E. (1989) Use of selectively trypsinized nucleosome core particles to analyze the role of the histone tails in the stabilization of the nucleosome. J. Mol. Biol. 206, 451– 463 8. Hays, J. J., Clark, D. J., and Wolffe, A. P. (1991) Histone contribution to the structure of DNA in a nucleosome. Proc. Natl. Acad. Sci. U. S. A. 88, 6829 – 6833 9. Imai, B. S., Yau, P. M., Baldwin, J. P., Ibel, K., May, R. P., and Bradbury, E. M. (1986.) Hyperacetylation of core histones does not cause unfolding of nucleosomes. Neutron scatter accords with disc structure of nucleosomes. J. Biol. Chem. 201, 8784 – 8792 10. Allfrey, V. G. (1977) in Chromatin and Chromatin Structure (Li, H. J., Eckhardt, R. A., eds) pp. 167–191, Academic Press, New York 11. Grunstein, M. (1997) Histone acetylation in chromatin structure and transcription. Nature (London) 389, 349 –352 12. Brownell, J. E., Allis, and C. D. (1996) Special HATs for special occasions: linking histone acetylation to chromatin assembly and gene activation. Curr. Opin. Genet. Dev. 6, 176 –184 13. Turner, M. T., and O’Neill, L. P. (1995) Histone acetylation in chromatin and chromosomes. Semin. Cell Biol. 6, 229 –236 14. Wade, P. A., Pruss, D., and Wolffe, A. P. (1997) Histone acetylation: chromatin in action. Trends Biochem. Sci 22, 128 –132 15. Kornberg, R. D., and Lorch, Y. (1999) Chromatin-modifying and remodeling complexes. Curr. Opin. Genet. Dev. 9, 148 –151 16. Clarke, D. J., O’Neill, L. P., and Turner, B. M. (1993) Selective use of H4 acetylation sites in the yeast S. cerevisiae. Biochemistry 294, 557–561 17. Borun, T. W., Pearson, D., and Paik, W. K. (1972) Studies of histone methylation during the HeLa S-3 cell cycle. J. Biol. Chem. 247, 4288 – 4298 18. Annunziato, A. T., Eason, M. B., and Perry, C. A. (1995) Relationship between methylation and acetylation of arginine rich histones in cycling and arrested HeLa cells. Biochemistry 34, 2916 –2924 19. Rice, J. C., and Allis, C. D. (2001) Histone methylation versus histone acetylation: new insights into epigenetic regulation. Curr. Opin. Cell Biol. 13, 263–273 20. Sobel, R. E., Cook, R. G., Perry, C. A., Annunziato, A. T., and Allis, C. D. (1995) Conservation of deposition-related acetylation sites in newly synthesized histones H3 and H4. Proc. Natl. Acad. Sci. U. S. A. 92, 1237–1241 21. Kaufman, P. D., Kobayashi, R., Kessler, N., and Stillman, B. (1995) The p150 and p60 subunits of chromatin assembly factor I: a molecular link between newly synthesized histones and DNA replication. Cell 81, 1105–1114 22. Brownell, J. E., Zhou, J., Ranalli, T., Kobayashi, R., Roth, S. Y., and Allis, C. D. (1996) Tetrahymena histone acetyltransferase A: a homolog to yeast Gcn5p linking histone acetylation to gene activation. Cell 84, 843– 851 23. Guarente L. (1995) Transcriptional coactivators in yeast and beyond. Trends Biochem. Sci. 20, 516 –521

508

Molecular & Cellular Proteomics 1.7

24. Hays, S. R., Dollard, C., Winston, F., Beck, S., Trowsdale, J., and Dawid, I. B. (1992) The bromodomain, a conserved sequence found in human, Drosophila and yeast proteins. Nucleic Acids Res. 20, 2603 25. Peterson, C. L, and Tamkun, J. W. (1995) The SWI/SNF complex: a chromatin remodeling machine? Trends Biochem. Sci 20, 143–146 26. Bone, R. J., Lavender, J., Richman, R., Palmer, M. J., Turner, B. M., and Kuroda, M. I. (1994) Acetylated histone H4 on the male X chromosome is associated with dosage compensation in Drosophila. Genes Dev. 8, 96 –104 27. Jeppesen, P., and Turner, B. M. (1993) The inactive X chromosome in female mammals is distinguished by a lack of histone H4 acetylation. A cytogenetic marker for gene expression. Cell 74, 281–289 28. Turner, B. M., Birley, A. J., and Lavender, J. (1992) Histone H4 isoforms acetylated at specific lysine residues define individual chromosomes and chromatin domains in Drosophila polytene nuclei. Cell 69, 375–384 29. Braunstein, M., Rose, A. B., Holmes, S. G., Allis, C. D., and Broach, J. R. (1996) Efficient transcriptional silencing in Saccharomyces cerevisiae requires heterochromatin acetylation pattern. Mol. Cell. Biol. 16, 4349 – 4356 30. Matthews, H. R. (1998) in Chromosomes and Chromatin (Adolf, K. W., ed) Vol. I, pp. 3–32, CRC Press, Inc., Boca Raton, FL 31. Allis, C. D., Chicoine, L. G., Richman, R., and Schulman, I. G. (1985) Deposition-related histone acetylation in micronuclei of conjugating Tetrahymena. Proc. Natl. Acad. Sci. U. S. A. 82, 8048 – 8052 32. Sobel, R. E., Cook, R. G., and Allis, C. D. (1994) Non-random acetylation of histone H4 by a cytoplasmic histone acetyltransferase as determined by novel methodology. J. Biol. Chem. 269, 18576 –18582 33. Couppez, M., Martin-Ponthieu, A., and Sautiere, P. (1987) Histone H4 from cuttlefish testis is sequentially acetylated. J. Biol. Chem. 262, 2854 –2860 34. Thorne, A. W., Kmiciek, D., Mitchelson, K., Sautiere, P., and CraneRobinson, C. (1990) Patterns of histone acetylation. Eur J Biochem. 193, 701–713 35. Waterborg, J. H. (1992) Identification of five sites of acetylation in alfalfa histone H4. Biochemistry 31, 6211– 6219 36. Turner, B. M. (1993) Decoding the nucleosome. Cell 75, 5– 8 37. Vidali, G., Botta, L. C., Bradbury, E. M, and Allfrey, V. G. (1978) Suppression of histone deacetylation leads to accumulation of multiacetylated forms of histone H3 and H4 and increased DNase I sensitivity of associated DNA sequences. Proc. Natl. Acad. Sci. U. S. A. 75, 2239 –2244 38. Edmonds, C. G., Loo, J. A., Smith, R. D., Fuciarelli, A. F., Thrall, B. D., Morris, J. G., and Springer, D. L. (1993) Evaluation of histone sequence and modifications by electrospray mass spectrometry and tandem mass spectrometry. J. Toxicol. Environ. Health 40, 159 39. Marvin, K. W., Yau, P., and Bradbury, E. M. (1990) Isolation and characterization of acetylated histone H3 and H4. J. Biol. Chem. 265, 19839 –19847 40. Clauser, K., Baker, P., and Burlingame, A. L. (1999) The role of accurate mass measurement (⫹/⫺ 10 ppm) in protein identification strategies employing mass spectrometry and database interrogation. Anal. Chem. 71, 2871–2882 41. Hong, L., Schroth, G. P., Matthews, H. R., Yau, P., and Bradbury, E. M. (1993) Studies of the DNA binding properties of histone H4 amino terminus. J. Biol. Chem. 268, 305–314 42. Puig, O. M., Belle´ s, E., Lo´ pez-Rodas, G., Sendra, R., and Tordera, V. (1998) Interaction between N-terminal domain of H4 and DNA is regulated by the acetylation degree. Biochim. Biophys. Acta 1397, 79 –90 43. Lee, D. Y., Hayes, J. J., Pruss, D., and Wolffe, A. P. (1993) A positive role for histone acetylation in transcription factor access to nucleosomal DNA. Cell 72, 73– 84 44. Pazin, M. J., and Kadonaga, J. T. (1997) What’s up and down with histone deacetylation and transcription. Cell 89, 325–328 45. Hansen, J. C., Tse, C., and Wolffe, A. P. (1998) Structure and function of core histone termini: more than meets the eyes. Biochemistry 37, 17637–17641 46. Vettese-Dadey, M., Grant, P. A., Hebbes, T. R., Crane-Robinson, C., Allis, C. D., and Workman, J. L. (1996) Acetylation of histone H4 plays a primary role in enhancing transcription factor binding to nucleosomal DNA in vitro. EMBO J. 15, 2508 –2518