ArticlePDF Available

Analysis of human acetylation stoichiometry defines mechanistic constraints on protein regulation

Authors:
  • The Novo Nordisk Foundation Center for Protein Research

Abstract and Figures

Lysine acetylation is a reversible posttranslational modification that occurs at thousands of sites on human proteins. However, the stoichiometry of acetylation remains poorly characterized, and is important for understanding acetylation-dependent mechanisms of protein regulation. Here we provide accurate, validated measurements of acetylation stoichiometry at 6829 sites on 2535 proteins in human cervical cancer (HeLa) cells. Most acetylation occurs at very low stoichiometry (median 0.02%), whereas high stoichiometry acetylation (>1%) occurs on nuclear proteins involved in gene transcription and on acetyltransferases. Analysis of acetylation copy numbers show that histones harbor the majority of acetylated lysine residues in human cells. Class I deacetylases target a greater proportion of high stoichiometry acetylation compared to SIRT1 and HDAC6. The acetyltransferases CBP and p300 catalyze a majority (65%) of high stoichiometry acetylation. This resource dataset provides valuable information for evaluating the impact of individual acetylation sites on protein function and for building accurate mechanistic models.
This content is subject to copyright. Terms and conditions apply.
ARTICLE
Analysis of human acetylation stoichiometry
denes mechanistic constraints on protein
regulation
Bogi Karbech Hansen1, Rajat Gupta1, Linda Baldus2,3, David Lyon4, Takeo Narita 1, Michael Lammers2,3,
Chunaram Choudhary 1& Brian T. Weinert1
Lysine acetylation is a reversible posttranslational modication that occurs at thousands of
sites on human proteins. However, the stoichiometry of acetylation remains poorly char-
acterized, and is important for understanding acetylation-dependent mechanisms of protein
regulation. Here we provide accurate, validated measurements of acetylation stoichiometry at
6829 sites on 2535 proteins in human cervical cancer (HeLa) cells. Most acetylation occurs
at very low stoichiometry (median 0.02%), whereas high stoichiometry acetylation (>1%)
occurs on nuclear proteins involved in gene transcription and on acetyltransferases. Analysis
of acetylation copy numbers show that histones harbor the majority of acetylated lysine
residues in human cells. Class I deacetylases target a greater proportion of high stoichiometry
acetylation compared to SIRT1 and HDAC6. The acetyltransferases CBP and p300 catalyze a
majority (65%) of high stoichiometry acetylation. This resource dataset provides valuable
information for evaluating the impact of individual acetylation sites on protein function and
for building accurate mechanistic models.
https://doi.org/10.1038/s41467-019-09024-0 OPEN
1Department of Proteomics, The Novo Nordisk Foundation Center for Protein Research, Faculty of Health and Medical Sciences, University of Copenhagen,
Blegdamsvej 3B, DK-2200 Copenhagen, Denmark. 2Institute of Biochemistry, Synthetic and Structural Biochemistry, University of Greifswald, Felix-
Hausdorff-Str. 4, Greifswald 17487, Germany. 3Institute for Genetics and Cologne Excellence Cluster on Cellular Stress Responses in Aging-Associated
Diseases, CECAD, University of Cologne, Joseph-Stelzmann-Str. 26, 50931 Cologne, Germany. 4Disease Systems Biology Program, The Novo
Nordisk Foundation Center for Protein Research, Faculty of Health and Medical Sciences, University of Copenhagen, Blegdamsvej 3B, DK-2200
Copenhagen, Denmark. Correspondence and requests for materials should be addressed to C.C. (email: chuna.choudhary@cpr.ku.dk)
or to B.T.W. (email: brian.weinert@gmail.com)
NATURE COMMUNICATIONS | (2019) 10:1055 |https://doi .org/10.1038/s41467-019-09024-0 | www.nature.com/naturecommunications 1
1234567890():,;
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Lysine N-ε-acetylation is a reversible protein posttranslational
modication (PTM) that was rst identied on histones1.In
the past decade, sensitive mass spectrometry (MS) techni-
ques enabled identication of thousands of acetylation sites on
diverse cellular proteins24. Acetylation can be enzymatically
catalyzed by lysine acetyltransferases, however, recent data indi-
cates that acetylation also arises from nonenzymatic reaction with
acetyl-CoA5,6. Nonenzymatic acetylation potentially targets any
solvent accessible lysine residue, suggesting that nonenzymatic
acetylation sites are likely to greatly outnumber acetyltransferase-
catalyzed sites. As a result, enzyme-catalyzed acetylation is easily
overlooked within a vast background of nonenzymatic acetyla-
tion, presenting a needle-in-a-haystack problem for identifying
these sites. Proteome-wide analyses of lysine acetylation should
focus on identifying parameters that will help prioritize the
functional relevance of individual sites and provide mechanistic
insights. These parameters include regulation by acetyl-
transferases and deacetylases, dynamic turnover rates, and the
stoichiometry of modication. Regardless of the origin of acet-
ylation, enzyme-catalyzed or nonenzymatic, understanding the
stoichiometry of modication is important for determining the
impact of acetylation on protein function and for building
accurate mechanistic models.
We developed a quantitative proteomics method to determine
acetylation stoichiometry at thousands of sites by measuring
differences in the abundance of native and chemically acetylated
peptides6,7. We subsequently rened our method by incorporat-
ing strict criteria for accurate quantication of acetylated pep-
tides8. However, the stoichiometry of acetylation in human cells
remains poorly characterized.
Here we determine acetylation stoichiometry at thousands of
sites in human cervical cancer (HeLa) cells. We validate our
results using known quantities of peptide standards, using
recombinant acetylated proteins, and by comparison with acety-
lated peptide intensity. This high-condence dataset is used to
calculate acetylation copy numbers in cells, to explore the rela-
tionship between stoichiometry and regulation by acetyl-
transferases and deacetylases, and to reveal mechanistic
constraints on protein regulation by acetylation.
Results
Measuring acetylation stoichiometry. We measured acetylation
stoichiometry in HeLa cells using partial chemical acetylation and
serial dilution SILAC (SD-SILAC) to ensure quantication
accuracy8(Fig. 1a). Two independent biological replicates were
performed, each using a different degree of chemical acetylation
and inverting the SILAC labeling between experiments. The
degree of chemical acetylation was estimated based on the median
reduction of unmodied peptides generated by tryptic cleavage at
one or two lysine residues (Supplementary Figure 1a). Based on
the estimated degree of chemical acetylation, we performed a
serial dilution of the chemically acetylated peptides to give
median ~1%, ~0.1%, and ~0.01% chemical acetylation. Acetylated
peptides were enriched and the differences between native
acetylated and chemically acetylated peptides quantied by MS
(Supplementary Data 1a). To ensure accurate quantication, we
required that the abundance of native acetylated peptides was
quantied by comparison with at least two different concentra-
tions of chemically acetylated peptides, and that the measured
SILAC ratios agreed with the serial dilution series. SILAC ratios
that did not follow the dilution series (allowing up to two-fold
variability) were dened as being inaccurately quantied, even
though one of the measurements may be correct. Quantication
error was reduced when the concentration of chemically acety-
lated peptides was most similar to native acetylated peptides
(Fig. 1b). However, quantication error was substantially higher
than in our previous experiments in bacteria8, likely due to the
greater complexity of the human proteome. The high error rates
highlight the need to control for quantication accuracy, and
show that comparing native acetylated peptides to just 1% che-
mically acetylated peptides results in a majority of false quanti-
cation (Fig. 1b). The measured stoichiometry of acetylated
peptides was signicantly and highly correlated between inde-
pendent experimental replicates (Fig. 1c). The precision of our
measurements was also highly reproducible; the median ratio of
stoichiometry between replicates was 0.95, and 90% of the mea-
surements varied by less than a factor of two between replicates
(Fig. 1d).
For high stoichiometry (>10%) acetylation, the difference
between native acetylated and chemically acetylated peptides
becomes too small to accurately measure by SILAC quantica-
tion, which is typically limited to differences greater than 2-fold
in magnitude. At 5% partial chemical acetylation, a peptide with
90% stoichiometry will have a SILAC ratio of 1.006, and a peptide
with 50% stoichiometry will have a SILAC ratio of 1.05
(Supplementary Figure 1b). These differences are too small to
accurately resolve, and can result in inaccurate stoichiometry
measurements that are out of bounds (greater than 100% or
negative). Due to the inherent limitations in calculating
stoichiometry for these peptides, we set a cutoff of maximum
10% stoichiometry, and classied all sites exceeding this cutoff as
having >10% stoichiometry. Only 16 peptides met these criteria,
and they harbored previously known high stoichiometry acetyla-
tion sites (Supplementary Data 1a), including; histones H3 K23
(2 different peptides), H3 K14, and H4 K16. Other peptides
harbored sites on acetyltransferases, such as CBP K1583
(2 different peptides) and N-alpha-acetyltransferase 50 K33, or
on proteins that catalyze reactions using acyl-CoAs, such as
hydroxymethylglutaryl-CoA lyase K48 and dihydroxyacetone
phosphate acyltransferase K643. These data show that, although
we are unable to accurately measure high stoichiometry
acetylation, we were able to identify these peptides, they represent
known or probable high stoichiometry acetylation sites, and they
constitute a small portion (0.2%) of the peptides analyzed.
Validating stoichiometry measurements. We used known
quantities of unmodied and acetylated peptide standards
(AQUA peptides) to determine stoichiometry directly at ten sites
on three proteins, cortactin (CTTN), nucleolin (NCL), and N-
acetyltransferase 10 (NAT10) (Supplementary Data 1b). Stoi-
chiometry determined by partial chemical acetylation (PCA) was
signicantly correlated (r=0.94) with stoichiometry determined
using AQUA peptides (Fig. 1e). Furthermore, stoichiometry
measurements differed by a factor of two or less for a majority
(7/10) of the analyzed peptides (Fig. 1f). Three sites showed 3.4-,
6-, and 7-fold higher stoichiometry by PCA, indicating over-
estimation of stoichiometry by PCA or underestimation by
AQUA. These differences occurred at the site-level and were
therefore not attributable to errors in protein quantication.
We think that the agreement between these two methods is
notable when considering all possible sources of variability in
each measurement.
We further validated our measurements using two recombi-
nant, site-specically acetylated proteins; malate dehydrogenase
(MDH2) K239ac and hydroxymethylglutaryl-CoA lyase
(HMGCL) K48ac. Recombinant acetylated proteins (SILAC-
light-labeled) were used as a spike-in standard to measure
acetylation stoichiometry in SILAC-heavy-labeled HeLa lysate.
Stoichiometry measured using two different concentrations of
recombinant acetylated protein (spike-in) agreed with our
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-09024-0
2NATURE COMMUNICATIONS | (2019) 10:1055 | https://doi.org/10.1038/s41467-019-09024-0 | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
measurements using PCA, further supporting the accuracy of our
stoichiometry dataset (Fig. 1g).
Stoichiometry measurements were additionally validated by
comparison to acetylated peptide intensity that was corrected for
differences in protein abundance. We previously showed that
abundance-corrected intensity (ACI) is correlated to acetylation
stoichiometry in yeast6. ACI was signicantly correlated with
acetylation stoichiometry in HeLa cells (Supplementary
a
b
eg
c
1%
10%
0.1%
0.01%
0.001%
1%
10%
0.1%
0.01%
0.001%
Log10 stoichiometry Exp.2
Log10 stoichiometry
Exp.1
1%
10%
0.1%
0.01%
0.001%
1%
10%
0.1%
0.01%
0.001%
Log10 stoichiometry
AQUA
Log10 stoichiometry
PCA
n = 3839, r = 0.89, P < 2e–16
n = 10, r = 0.94, P = 7e–5
0%
10%
20%
30%
40%
50%
60%
70%
80%
90%
1%
0.1%
0.01%
1%
0.1%
0.01%
Exp.1 Exp.2
Median % chemically acetylated peptides
Quantification error
Count
−3 −2 −1 0
Log2 stoichiometry
Exp.1/Exp.2
250
200
150
100
50
0
SILAC light (L )
Native acetylation (-nAc)
-nAc
SILAC heavy (H )
Chemical acetylation (-cAc)
-cAc
-cAc
cAc-
cAc-
cAc-
-cAc
-cAc
-cAc
Combine and
digest to peptides
Analyze data and
calculate stoichiometry
-nAc
-cAc
cAc-
-cAc
-cAc
-nAc
-cAc
cAc-
-cAc
-cAc
Antibody enrich
-Ac peptides
Quantify by MS
and serial dilution
1:1 dilution
m/z
1:10 dilution
m/z
1:100 dilution
m/z
SILAC
ratio H/L
= 7.0
SILAC
ratio H/L
= 0.7
SILAC
ratio H/L
= 0.07
2. Filter for accurate quantitation
3. Calculate stoichiometry
SILAC ratio ~ dilution series
MaxQuant
1. Computational analysis
S = stoichiometry
R = SILAC ratio cAc/nAc
C = % chemical Ac
S = C / (R - (1-C))
Experiment 1: SILAC L = Native acetylation; SILAC-H = Chemical acetylation; % chemical acetylation (C) = 3.5%
Experiment 2: SILAC L = Chemical acetylation; SILAC-H = Native acetylation; % chemical acetylation (C) = 10.4%
3
2
1
0
−1
−2
−3
Log2 ratio stoichiometry
PCA/AQUA
MDH2 K239ac
HMGCL K48ac
0.01 %
0.1%
Native
Native
Intensity
m/z m/z
Spike-in stoichiometry
0.015%
0.018%
PCA stoichiometry
1%
10%
Native
m/z m/z
Native
Intensity
Spike-in stoichiometry
7.8%
>10%
PCA stoichiometry
90%
321
f
d
Fig. 1 Measuring acetylation stoichiometry. aDiagram of the method used to measure acetylation (Ac) stoichiometry. bThe degree of quantication error
as determined by the fraction of SILAC ratios at each concentration of chemically acetylated peptides that was not consistent with SILAC ratios measured
in at least one different concentration of chemically acetylated peptides. cThe correlation between stoichiometry measured in independent experimental
replicates. The number of peptides (n), Pearsons correlation (r), and P-value (P) of correlation are shown. dLow absolute variability between experimental
replicates. The histogram shows the distribution of Log2 ratios of stoichiometry in Experiment 1/Experiment 2 (Exp.1/Exp.2). eThe correlation between
stoichiometry measured using partial chemical acetylation (PCA) and absolute quantication (AQUA) peptide standards. fLow absolute variability
between stoichiometry measurements made by PCA and AQUA. gValidation of stoichiometry measurements using recombinant acetylated (100%)
proteins as a spike-in standard. Stoichiometry was measured at two different concentrations of spike-in protein (SILAC light, red) compared to SILAC
heavy-labeled HeLa (blue) for each acetylation site. Source data are provided as a Source Data le
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-09024-0 ARTICLE
NATURE COMMUNICATIONS | (2019) 10:1055 | https://doi.org/10.1038/s41467-019-09024-0 | www.nature.com/naturecommunications 3
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Figure 1c), however, the predictive power of this correlation was
modest (r=0.480.52). There are several reasons for this modest
correlation. Firstly, peptide intensity is inherently variable.
Secondly, protein abundance estimates may be inaccurate. We
found that outlier data points using iBAQ-based protein
abundance were not outliers when using copy-number-based
protein abundance (Supplementary Figure 1c), indicating that
variability in protein abundance measurements contributes to
disagreement between ACI and stoichiometry measurements.
Thirdly, antibody-based acetylated peptide enrichment may be
peptide-sequence biased, which will introduce further variability.
Regardless of these limitations, ACI provides an easy method to
estimate the relative stoichiometry of acetylation sites, and the
signicant correlation with our stoichiometry measurements by
PCA provides further support for the accuracy of our
measurements.
Copy number limits the detection of acetylated peptides. The
detection of acetylated peptides is biased to abundant proteins
(Fig. 2a). Furthermore, the fraction of lysines that are detected as
acetylated on any given protein is signicantly correlated with
protein abundance (Fig. 2b). This bias is found in every acetylome
dataset that we have examined (Supplementary Figure 2a)912,
and indicates that acetylation occurs on most lysine residues in
cells and that protein abundance is a limiting factor in the
detection of acetylated peptides. These data further support the
notion that all solvent accessible lysine residues are acetylated to
some degree, either enzymatically or nonenzymatically.
Deep proteome measurements detect unmodied peptides
from proteins whose abundance spans seven orders of magnitude
(Fig. 2a)13. This raises the question of why we detect so few
acetylated peptides without antibody enrichment (Fig. 2c). The
signal intensity of acetylated peptides is comparable to unmodi-
ed peptides in the mass spectrometer (Supplementary Figure 2b,
c), indicating that we should be able to detect acetylated peptides
as readily as unmodied peptides. We compared copy numbers
for unmodied peptides to copy numbers for acetylated peptides
as determined from our stoichiometry measurements. The
distribution of acetylated peptide copy number shows that, in
the absence of acetylated peptide enrichment, most acetylated
peptides are at or below the detection limit of the mass
spectrometer, even in deep proteome measurements (Fig. 2d).
In contrast, acetylated peptides that were detected without
antibody enrichment occurred at copy numbers that were within
the detectable range of unmodied peptides (Fig. 2e). Thus, our
stoichiometry measurements are consistent with the inability to
detect acetylated peptides without enrichment. Strikingly, our
data indicate that some acetylation events are so rare that they
occur at a copy number that is less than one per cell (Fig. 2d).
Properties of high stoichiometry acetylation. We measured the
stoichiometry of acetylated peptides; however, individual acet-
ylation sites may occur on multiple different peptides due to
incomplete tryptic digestion, protein N-terminal acetylation, or
oxidized methionine residues. To examine acetylation stoichio-
metry at the site-level, we calculated the summed stoichiometry of
peptides containing the same acetylation site (Supplementary
Data 1c). This resulted in stoichiometry measurements for
6829 sites, with a median stoichiometry of just 0.02% (1/4000
molecules) (Fig. 3a). This represents very low levels of acetylation
for most sites, only 1% (66 sites) displayed stoichiometry >1%,
and ~15% (1014 sites) displayed stoichiometry >0.1%.
We performed UniProt keyword enrichment analysis to
examine the functional categories of proteins that are associated
with higher stoichiometry (>0.23% or >1%) acetylation (Fig. 3b).
Higher stoichiometry acetylation was overrepresented on nuclear
proteins involved in chromatin regulation and transcription. This
observation is consistent with the known nuclear functions of
acetyltransferases, deacetylases, and acetylated lysine-binding
bromodomain proteins. In fact, the keywords Bromodomain
and Acetyltransferase were signicantly enriched in the group of
proteins with high stoichiometry acetylation. We were unable to
calculate stoichiometry for doubly acetylated peptides because of
the low frequency of chemical acetylation at both positions.
08
0
4
0
300
200
100
Log10 copy number Log10 copy number Log10 copy number
Log10 copy number
Count (proteins)
Count
(peptides)
Proteins (8425)
Ac proteins (2488)
ad
e
c
4325
100%
10%
1%
Log10 % lysines acetylated
on each protein
2488 Ac proteins
6753 Ac sites
r = 0.65, P < 2e−16
This study
Bekker-Jensen et al.
Proteins
11,264
10,118
Peptides
398,893
170,251
Ac peptides
732
282
% Ac peptides
0.18%
0.17%
Detection of acetylated peptides without antibody enrichment
−2 0
0
6000
5000
4000
3000
2000
1000
Count (peptides)
Peptides (158,796)
Ac peptides (7479)
246 2468
202468
b
67
Fig. 2 Stoichiometry limits the detection of acetylated peptides. aAcetylation is biased to detection on abundant proteins. Protein copy number estimates
are from42.bThe fraction of acetylated lysines detected on any given protein is correlated with protein abundance. The scatterplot shows the %
lysines acetylated and copy numbers of 2488 acetylated proteins containing 6753 acetylation sites. The Pearsons correlation (r), and P-value (P)of
correlation are shown. cThe number of peptides and acetylated peptides (Ac peptides) detected in deep proteome measurements from this study and13.
dThe distribution of peptide copy numbers from a deep proteome measurement and acetylated peptide copy numbers calculated from the peptide
stoichiometry and protein copy number. eThe distribution of acetylated peptide copy numbers for acetylated peptides that were detected without prior
antibody enrichment. Source data are provided as a Source Data le
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-09024-0
4NATURE COMMUNICATIONS | (2019) 10:1055 | https://doi.org/10.1038/s41467-019-09024-0 | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
However, given the low stoichiometry of acetylation, doubly
acetylated peptides are unlikely to occur by random chance and
most likely reect the activity of acetyltransferases. Consistent
with this idea, doubly acetylated peptides occurred on proteins
that were overrepresented for the same UniProt keywords that
were associated with high stoichiometry acetylation (Fig. 3b).
Thus, sites occurring on doubly acetylated peptides may occur at
high stoichiometry and are likely to be enzyme-catalyzed.
To investigate the relationship between stoichiometry and
subcellular compartmentalization we used immunouorescence-
based protein localization as determined by the Human Protein
Atlas14. Acetylation stoichiometry was broadly distributed and
mostly similar in every subcellular compartment analyzed
(Fig. 3c). Mitochondrial acetylation occurred at a slightly, yet
signicantly (P<5e
5, Wilcoxon test), higher median stoichio-
metry. However, mitochondria contained the smallest fraction of
high (>1%) stoichiometry acetylation sites. In contrast, the
nucleus contained the greatest fraction of high stoichiometry
sites, which was approximately an order of magnitude greater
than in mitochondria (Fig. 3c).
We used IceLogo15 to determine whether high stoichiometry
acetylation was associated with neighboring amino acids.
Cysteine residues were notably overrepresented for sites with
>0.23% stoichiometry (10-fold higher than median stoichiome-
try), particularly in the 4, 3, and 2 positions (Fig. 3d).
However, this bias was absent when examining sites with >1%
stoichiometry, indicating that this overrepresentation was asso-
ciated with sites with moderately elevated stoichiometry.
Remarkably, sites with cysteine residues in the 4, 3, or 2
position constituted 35% (159/460) of the sites with >0.23%
stoichiometry. UniProt keyword enrichment analysis of the
proteins harboring these sites found a variety of enriched
keywords (Supplementary Data 1d). However, unlike high
stoichiometry acetylation in general (Fig. 3b), keywords describ-
ing processes associated with nuclear acetyltransferases, such as
Nucleus, Transcription, and Chromosome were notably absent.
These data suggest that cysteine residues may promote none-
nzymatic acetylation of downstream lysine residues, and these
sites constitute a substantial portion (35%) of sites with an
elevated (>0.23%) stoichiometry of acetylation. This conclusion is
a
0
400
300
200
100
Count (sites)
10%
100%
1%
0.1%
0.01%
0.001%
Stoichiometry
>0.23% stoichiometry
>0.23% stoichiometry
>1% stoichiometry
>1% stoichiometry
Doubly acetylated peptides
% associated with keyword
>2-fold keword enrichment
c
Cytosol
Nucleus or
nucleoplasm
Mitochondria
Vesicles
Plasma
membrane
n 2745 2433 758 463 741
Median – 0.022% 0.025% 0.032% 0.023% 0.020%
>1% – 0.77% 1.40% 0.13% 1.08% 0.54%
0.001%
0.01%
0.1%
1%
10%
100%
Stoichiometry
25
–25
12
–12
Fold change
(P value < 0.05)
25
–25
12
–12
Fold change
(P value < 0.05)
d
–15 –13–12 –11 –9 –8 –7 –6 –5 –4 –3 –2 –1 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15–10–14
–15 –13–12 –11 –9 –8 –7 –6 –5 –4 –3 –2 –1 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15–10–14
b
Transcription regulation
Nucleosome core
Nucleotide−binding
Transferase
Acyltransferase
Isopeptide bond
Transcription
Methylation
Chromosome
RNA−binding
Nucleus
Ubl conjugation
Helicase
Acetylation
Chromatin regulator
DNA−binding Thiamine pyrophosphate
ATP−binding
Lyase
Bromodomain
Chromosome
Nucleus
Nucleosome core
Peroxisome
Nucleotide−binding
Acetylation
Chromatin regulator
Methylation
Helicase
Isopeptide bond
Bromodomain
Transcription regulation
Ubl conjugation
Citrullination
Activator
DNA−binding
Transcription
Coiled coil
Acyltransferase
Nucleus
Activator
Citrullination
Nucleosome core
Chaperone
DNA−binding
Methylation
Ribonucleoprotein
Glycolysis
Acetylation
Ubl conjugation
Isopeptide bond
Chromosome
Chromosomal rearrangement
Acyltransferase
Bromodomain
Coiled coil
50% 100%
10% 75%
25%
Fig. 3 Properties of high stoichiometry acetylation. aThe distribution of acetylation site stoichiometry for the 6829 sites measured in this study. bUniProt
keyword enrichment for the indicated classes of high stoichiometry acetylation sites (>0.23% and >0.1%) and for doubly acetylated peptides. The sizeof
the text is related to the fraction of sites associated with the keyword, and keywords that were more than two-fold enriched are colored red. cSubcellular
compartment analysis based on the Human Protein Atlas14. Category scatterplots show the distributions of acetylation site stoichiometry in each
subcellular compartment. The number (n) of sites analyzed, median stoichiometry (median), and percentage of sites with >1% stoichiometry (>1%) is
shown. dAmino acid sequence logos of the indicated classes of acetylation sites using IceLogo15. Source data are provided as a Source Data le
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-09024-0 ARTICLE
NATURE COMMUNICATIONS | (2019) 10:1055 | https://doi.org/10.1038/s41467-019-09024-0 | www.nature.com/naturecommunications 5
Content courtesy of Springer Nature, terms of use apply. Rights reserved
supported by a recent study that uncovered a similar bias for
higher stoichiometry acetylation at sites with proximal cysteine
residues16.
Histone acetylation. We measured stoichiometry at 57 histone
sites and found that high stoichiometry acetylation was mostly
restricted to sites on the N-terminal tails of core histones H2B,
H3, and H4 (sites on H2A were not measured) (Fig. 4a). The
stoichiometry of histone H3 and H4 acetylation sites has been
extensively studied1720, and our measurements are comparable
to these previous measurements (Fig. 4b). However, our method
does not measure the stoichiometry of doubly acetylated peptides
that can arise from the lysine-rich N-terminal tails of the core
histones H2A, H2B, H3, and H4. Feller et al.18 found that the
stoichiometry of mono-acetylated histones H3 and H4 were more
abundant than di-acetylated H3 and H4, with the exception of H4
K5 +K12 and H3 K18 +23 (which is more abundant than K18
alone, but less abundant than K23 alone). We detected doubly
acetylated peptides containing the following sites on H2A (K5 +
K9, K9 +K11, K11 +K13), H2B (K5 +K11, K11 +K12, K15 +
K16, K16 +K20, K20 +K23, K34 +K43, and K116 +K120), H3
(K9 +K14, K18 +K23, and K27 +K36), and H4 (K8 +K12 and
K12 +K16). The stoichiometry of di-acetylated peptides remains
unexplored for H2B, and these peptides may be more abundant
than their mono-acetylated counterparts used to calculate stoi-
chiometry in this study. In addition, some acetylated peptides
from histone tails may not be detected because of their small size.
Thus, our estimates of histone acetylation stoichiometry may
underestimate the actual native stoichiometry at these positions
because these sites also occur on di-acetylated peptides or pep-
tides that we are unable to detect with our methodology.
Regardless, our data suggest that the stoichiometry of H2B
acetylation ranges from 0.5% to 5.6%. N-terminal H2B sites are
primarily acetylated by the CBP/p300 acetyltransferases12, which
also target H3K27 and H3K36. N-terminal H2B sites (K5, K11,
Non-histone
Histone (this study)
Histone (other)
H3K9
H3K14
H3K18
H3K23
H3K27
H3K36
H4K5
H4K8
H4K12
H4K16
0%6.
23.0%
3.0%
3.0%
6.0%
30.0%
0.9%
21.0%
10.7%
55.5%
0.3%
0.3%
9.8%
6.0%
14.9%
5.5%
5.3%
17.6%
8.9%
7.1%
13.3%
11.0%
27.8%
4.1%
25.1%
2.7%
6.9%
12.7%
35.7%
Histone
Ac site
Abshiru et al
>15%
>20%
>10%
This study
Feller et al
Zhou et al
Zheng et al
Ac copy number per cell
5e7
4e7
3e7
2e7
1e7
0
N-terminal
Histone Ac site Stoichiometry
H3.1 23 >20%
H3.1 14 >15%
H4 16 >10%
H2B 2-F 20
H2B 1-K 20
H1.4 17
5
5
5
5
5
5
2.1%
H2B 1-K 1.6%
H2B 1-M 20
H2B 1-M 0.98%
H2B 1-B 20 0.66%
H1.2 17 0.59%
H2B 1-D 0.43%
H2B 1-H 0.29%
H2B 1-L 0.26%
H2B 3-B 0.20%
H1.2 21 0.17%
H1.1 17 0.14%
C-terminal
Histone Ac site Stoichiometry
H2A.1 291 0.099%
H1.5 35 0.067%
H1.5 17 0.064%
H1.0 12 0.062%
H2A.1 303 0.061%
0.059%H1.2
H1.4 46 0.049%
H1.5 66 0.048%
0.039%H1.4 106
H1.4 63 0.036%
H1.5 49 0.034%
H1x 143 0.031%
H1.1 182 0.030%
H1.4 90 0.029%
H4 91 0.029%
H1.0 27 0.024%
H4 77 0.023%
H2A.1 115 0.022%
H2A.Z 115 0.021%
H2A 1-C 95
C-terminal
Histone Ac site Stoichiometry
H1.4 117
H2B 1-O 108 0.017%
0.017%
0.019%
H2B 1-O 120
H1.0 59 0.016%
0.015%
0.012%
0.012%
0.011%
0.011%
0.011%
0.010%
0.010%
0.009%
0.009%
0.009%
0.009%
0.009%
H2B 1-O 46
H2A 1 95
H1x 115
H2B 1-D 120
H2B 1-O 116
H1.5 78
H2B F-S 85
H2B 1-D 43
H2A 2-A 95
H1.0 52
H2B 1-K 120
H1.4 75
H1x 90
H1.0 69 0.008%
H1.0 40 0.007%
H1.0 82 0.006%
188
98
62
49
38
28
17
14
6
Short
Long
Exposure
H3 (15 kD)
H2A/H2B (14 kD)
H4 (11 kD)
26%
74%
16%
46%
38%
6e7
5.6%
2.9%
1.2%
168
0.019%
a
bcd
Fig. 4 Stoichiometry of histone acetylation. aThe diagram shows all histone acetylation sites whose stoichiometry was determined in this study. The sites
are ordered by descending stoichiometry. Note that high stoichiometry sites occur on the N-termini of core histones. bThe stoichiometry of histone
acetylation sites as determined in four independent studies1720.cAn anti-acetylated lysine immunoblot of HeLa whole cell lysate. Cells were boiled in 2%
LDS to ensure histone extraction. Histones are annotated based on their expected molecular weight. dHistone acetylation sites constitute a majority of
acetylated lysine residues in cells. Stoichiometry and protein copy numbers were used to calculate the number of acetylated lysine residues for the
indicated classes of proteins. Source data are provided as a Source Data le
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-09024-0
6NATURE COMMUNICATIONS | (2019) 10:1055 | https://doi.org/10.1038/s41467-019-09024-0 | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
K12, K15, and K16) show faster deacetylation kinetics compared
to K2012. Interestingly, the stoichiometry of K20 is greater than
K5, indicating that the lower stoichiometry at K5 is possibly due
to its more rapid turnover.
Histones are some of the most abundant and most highly
acetylated proteins in cells. Anti-acetylated lysine antibodies
prominently detect histone acetylation in western blots of whole
cell lysates, suggesting that histones harbor most of the acetylated
lysine residues in cells (Fig. 4c). Using our acetylated lysine copy
number estimates we found that histone acetylation accounted for
74% of the acetylated lysine residues in cells (Fig. 4d). If we
include histone sites whose stoichiometry was measured by
independent studies1720, the fraction of acetylated lysines
occurring on histones increases to 84% (Fig. 4d). These estimates
do not account for several sites on H2B (K11, K12, K15, K16, and
K23), as well as sites on H2A (K4, K5, K7, K9, K11, and K13).
Thus, histone acetylation likely accounts for an even greater
proportion of the acetylated lysine residues found in cells.
Regulation by deacetylases. We analyzed the stoichiometry of
lysine deacetylase (KDAC)-regulated acetylation sites by com-
paring the data obtained in this study with a previous analysis of
deacetylase inhibitors in HeLa cells21. Class I KDACs (HDAC 1,
2, 3, and 8) are specically targeted by the class I inhibitors
apicidin, MS-275, valproic acid, and sodium butyrate; the class IIb
KDAC HDAC6 is specically targeted by tubacin; and nicotina-
mide inhibits the activity of NAD+-dependent Sirtuin deacety-
lases, but mostly affected SIRT1-regulated sites in mammalian
cells21. To analyze class I KDAC regulated sites, we used the
median increased acetylation caused by apicidin, MS-275, val-
proic acid, and sodium butyrate.
KDAC inhibitors regulated sites with a broad range of
stoichiometry (Fig. 5a). Class I KDAC inhibitors regulated
substantially greater proportions of moderately increased
(>0.23%) and high stoichiometry (>1%) acetylation compared
to tubacin or nicotinamide. The stoichiometry of class I regulated
sites was signicantly higher than not-regulated (NR) sites, while
the distributions of tubacin and nicotinamide regulated sites was
not signicantly different than NR sites. Furthermore, sites that
were most sensitive to KDAC inhibitors (>4× increased
acetylation) showed an increasing proportion of higher stoichio-
metry sites for the class I inhibitors, but stayed the same or
decreased for tubacin and nicotinamide (Fig. 5a). Thus, while
tubacin and nicotinamide regulate a greater portion of the
acetylation sites quantied in this study (8.6% and 8.8%,
respectively) than class I inhibitors (2.6%), the class I inhibitors
regulate a greater proportion of higher stoichiometry acetylation.
Regulation by the CBP and p300 acetyltransferases. The
homologous Creb-binding protein (CBP)/E1A-binding protein
p300 (p300) acetyltransferases are important regulators of cell-
type-specic and signaling-regulated gene expression22. CBP/
p300 acetyltransferase activity is essential for promoting gene
transcription and CBP/p300 targets a large proportion of the
acetylome12. CBP/p300-regulated sites constituted 12.7% of the
sites analyzed in this study, indicating that CBP/p300 targets
more than one out of every ten acetylation sites. CBP/p300 tar-
geted a similar proportion (11.5%) of low stoichiometry ( < 0.2%)
sites, and an increasing proportion of higher stoichiometry sites,
up to 65% of the sites with stoichiometry exceeding 1% (Fig. 5b).
Thus, CBP/p300 acetylates a majority of high (>1%) stoichio-
metry acetylation sites. The stoichiometry of CBP/p300-regulated
sites also increased with the degree of downregulated acetylation
in the absence of CBP/p300 catalytic activity (Fig. 5c), indicating
that the sites most affected by loss of CBP/p300 tend to be more
highly acetylated.
Stoichiometry of functionally characterized sites. The activity
and subcellular localization of eukaryotic translation initiation
factor 5 A (eIF5A) is regulated by PCAF-catalyzed acetylation at
K4723. We found that eIF5A was more than 10% acetylated at
K47, consistent with a regulatory role for acetylation at this
position. DNA methyltransferase 1 (DNMT1) is acetylated by the
TIP60 acetyltransferase, and acetylation promotes ubiquitin-
0.001%
0.01%
0.1%
1%
10%
100%
Stoichiometry
0.001%
0.01%
0.1%
1%
10%
100%
Stoichiometry
2–4x reduced
4–8x reduced
>8x reduced
Unregulated
>2x
>4x
NR
>2x
>4x
NR
>2x
>4x
NR
Class I KDAC
inhibitors
Tubacin Nicotinamide
2.1%
16.7%
20.0%
3.6%
2.7%
2.7%
2.7%
5.6%
2.7%
9.3%
38.1%
40.0%
13.2%
8.9%
9.1%
10.5%
12.7%
8.1%
Fraction >1%
Fraction >0.23%
P = 2.5e–7 N.S. N.S.
0%
10%
20%
30%
40%
50%
60%
70%
All sites
<0.2%
>0.2%
>0.5%
>1%
Fraction regulated by CBP/p300
Stoichiometry
abc
Fig. 5 Stoichiometry of deacetylase- and CBP/p300-regulated acetylation sites. aThe category scatterplot shows the distributions of acetylation sites that
are not regulated (NR), more than two-fold (>2×) upregulated, or more than four-fold (>4×) upregulated, by the indicated deacetylase inhibitors as
determined by21. Class I KDAC inhibitors primarily target HDACs 1, 2, 3, and 8, and were determined by the median SILAC ratio of apicidin, MS-275,
valproic acid, and sodium butyrate-treated HeLa cells. Tubacin is an HDAC6 inhibitor and nicotinamide inhibits Sirtuin deacetylases, but the regulated sites
are mostly attributed to SIRT121.bCBP/p300 regulates an increasing fraction of high stoichiometry acetylation sites. CBP/p300-regulated sites were
determined by12.cAcetylation sites that are most affected (>8× reduced) by loss of CBP/p300 activity have higher median stoichiometry than sites that
are only modestly affected (2 reduced). Source data are provided as a Source Data le
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-09024-0 ARTICLE
NATURE COMMUNICATIONS | (2019) 10:1055 | https://doi.org/10.1038/s41467-019-09024-0 | www.nature.com/naturecommunications 7
Content courtesy of Springer Nature, terms of use apply. Rights reserved
dependent protein turnover24. We found that DNMT1 harbored
high stoichiometry acetylation at K335 (0.57%) and K675 (0.1%).
Pyruvate dehydrogenase E1 alpha 1 subunit (PDHA1) is acety-
lated at K321 by acetyl-CoA acetyltransferase 1 (ACAT1), which
recruits PDH kinase (PDK) to inhibit pyruvate dehydrogenase
activity25. We found that PDHA K321 was 0.6% acetylated,
supporting high stoichiometry acetylation at this position.
Glucose-6-phosphate dehydrogenase (G6PD) is reported to be
negatively regulated by KAT9-dependent acetylation of K403, and
activated by SIRT1-dependent deacetylation26. We found that
G6PD is just 0.02% acetylated at K403. Acetylation of
phosphoglycerate kinase 1 (PGK1) at K220 disrupts its activity
by inhibiting binding to ADP27. We found that PGK1 is 0.03%
acetylated at K220 in HeLa cells. PCAF is reported to regulate
cyclin dependent kinase 2 (CDK2) activity by acetylating CDK2
at K33 in the ATP-binding active site28. We found that CDK2
K33 is just 0.05% acetylated, however, CDK1 K33 was 4.5%
acetylated, suggesting that acetylation may play a greater role in
regulating CDK1 activity3. In each of the above examples,
acetylation is reported to reduce enzymatic activity, however, the
stoichiometry of acetylation suggests that acetylation would need
to be dramatically increased (100-fold or more) in order to have a
substantial impact on protein function.
Low stoichiometry acetylation does not necessarily indicate a
lack of function. Rather, the mechanism-of-action determines
whether low stoichiometry acetylation is sufcient to regulate
protein function. Acetylation that imparts a gain-of-function or
regulates protein activity at a specic time and/or place, could be
regulated by low stoichiometry acetylation. For example, acetyla-
tion at K220 regulates the activity of microtubule associated
protein RP/EB family member 1 (EB1), specically during mitosis
and at spindle microtubule plus ends29. Although we found that
EB1 K220 was only 0.02% acetylated, the mechanism-of-action is
consistent with the low observed stoichiometry of modication.
Discussion
Here we provide a validated resource of acetylation stoichiometry
at thousands of sites in the most widely studied human cell line
(HeLa). Our data shows that the vast majority of acetylation
occurs at very low stoichiometry. Thus, as a general rule, the
mechanisms by which acetylation regulates protein function
should agree with a low stoichiometry of modication. There are
a large number of metabolic enzymes whose catalytic activity is
reduced by site-specic acetylation30. Many studies relied on
acetylation mimicking glutamine-substitution mutations to assay
the impact of acetylation on these proteins, a method that results
in 100% stoichiometry of modication. However, our measure-
ments indicate that the vast majority of acetylation occurs at a
stoichiometry that is much less than 1%, and is therefore not
likely to impact protein activity through a loss-of-function
mechanism at a single acetylation site. Low stoichiometry acet-
ylation is compatible with gain-of-function mechanisms, or in
processes that are spatially or temporally restricted. However,
even for gain-of-function mechanisms such as increased catalytic
activity, a high degree of acetylation may be required to have a
measurable impact on overall activity. Thus, understanding the
stoichiometry of acetylation is important for formulating accurate
mechanistic models when evaluating the impact of acetylation on
protein function. High stoichiometry sites may be particularly
interesting because they are more likely to be enzyme-catalyzed.
However, more studies are required to determine if high stoi-
chiometry is a good indicator of functional importance. Enzy-
matic acetylation does not necessarily indicate a regulatory
function, and high stoichiometry acetylation may also occur by
nonenzymatic mechanisms.
We previously showed that the Sirtuin deacetylases SIRT3 and
CobB suppress acetylation at hundreds of sites to levels that are
equal to or less than the median stoichiometry of modication7,8.
These data support the idea that Sirtuin deacetylases may have a
general role suppressing nonenzymatic acetylation to preserve
protein function31. Here we nd that SIRT1 and HDAC6 also
suppress acetylation at regulated sites to levels that are compar-
able to the median stoichiometry of acetylation. SIRT1 likely
suppresses the activity of nuclear acetyltransferases, ensuring
tight control over the sites targeted by these enzymes. HDAC6
deacetylates a large number of cytoplasmic acetylation sites, most
of which are acetylated by unidentied acetyltransferases, or
nonenzymatically. Interestingly, the HDAC6 inhibitor Bufexamac
sensitizes cells to nonenzymatic acetylation caused by aspirin32,
suggesting that HDAC6 may have a role in protecting cells from
nonenzymatic acetylation stress. It is important to note that the
activity of these deacetylases as general suppressors of acetylation
at hundreds of sites, or as dynamic regulators of individual
acetylation sites, is not mutually exclusive. Deacetylases may also
target hundreds of acetylation sites for no reason whatsoever; if
such activity is not evolutionarily disadvantageous, it will not be
selected against.
Our data highlights the outsized impact of CBP and p300 on
the acetylome. CBP/p300 targets up to 20% of all acetylation sites
in cells, and acetylates proximal proteins in a sequence-
independent manner12. Here we nd that CBP/p300 targets a
majority (65%) of high (>1%) stoichiometry acetylation. How-
ever, CBP/p300 also acetylated many sites to a low stoichiometry
of modication. These low stoichiometry sites likely include both
functionally important regulatory acetylation and non-functional,
off-target acetylation. This presents a challenge to prioritizing
sites for mechanistic analyses and suggests that additional para-
meters, such as dynamic turnover rates12 and conditional reg-
ulation, are needed to identify potentially interesting sites.
Several other groups have also used chemical acetylation to
measure acetylation stoichiometry, often reporting substantially
higher stoichiometry than we observed in our experiments19,33,34.
One crucial difference between these studies and our own is that
we use a low degree of chemical acetylation (~510%) and dilute
the fraction of chemically acetylated peptides to allow for accurate
quantication8. Other studies used stable-isotope-labeled acet-
ylating agents, such acetic anhydride and NHS-acetate, to com-
pletely acetylate all free lysine residues and isotopically label the
chemically acetylated fraction at the same time19,33,34. While this
is an elegant approach, incomplete isotopic labeling of the acet-
ylating agents limits the resolution to stoichiometry greater than
12%, and it is not possible to independently dilute the chemi-
cally acetylated peptides to ensure accurate quantication. Most
of these studies19,3335 did not validate their results using
orthogonal methods, and the accuracy of their measurements is
likely impacted by the limited dynamic range of accurate quan-
tication by mass spectrometry36,37. One study found that ~75%
of their measurements were impacted by false quantication19,
and we found that ~90% of our measurements were inaccurate
when quantifying the differences between completely (100%)
acetylated peptides and native acetylated peptides8. The differ-
ences between 100% acetylated and native acetylated peptides are
likely too large to be accurately quantied by MS. Given the
differences in acetylation stoichiometry reported by independent
studies, care should be taken to carefully validate these
measurements.
The paucity of high stoichiometry acetylation detected in our
study is somewhat disappointing. However, as we show, it is
consistent with the inability to detect acetylated peptides in deep
proteome measurements. One possibility is that proteins are
deacetylated during protein extraction, resulting in uniformly
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-09024-0
8NATURE COMMUNICATIONS | (2019) 10:1055 | https://doi.org/10.1038/s41467-019-09024-0 | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
decreased acetylation. We consider this unlikely for several rea-
sons. We and others can quantify acetylation changes that occur
in vivo, indicating that these differences are preserved during
protein extraction. We furthermore compared acetylation after
lysing cells by three different methods and found that acetylation
levels were unaffected by the method of protein extraction
(Supplementary Figure 3). Thus, we believe that our measure-
ments indicate the stoichiometry of acetylation that occurs on
proteins inside cells.
Methods
Cell culture. HeLa (ATCC: CCL-2) cells were tested for mycoplasma con-
tamination and grown in DMEM supplemented with 10% FBS, 2 mM L-glutamine,
and 1% penicillin/streptomycin. SILAC media was supplemented with arginine and
lysine (SILAC Light) or with heavy isotope-labeled arginine (13C
6,
15N
4
-arginine,
Sigma) and lysine (13C
6,
15N
2
-lysine, Cambridge Isotope Laboratories) (SILAC
heavy) in media containing dialyzed serum (Sigma). Cells were cultured at 37 °C in
a humidied incubator at 5% CO
2
. At a conuency of ~90%, cells were washed
twice with PBS and lysed in ice-cold modied RIPA buffer (50 mM Tris, pH 7.5,
150 mM NaCl, 1 mM EDTA, 1x mini complete protease inhibitor cocktail (Roche),
10 mM nicotinamide, and 5 μM trichostatin A). Lysates were mixed with 1/10
volume of 5 M NaCl to release chromatin-bound proteins and incubated for 15 min
on ice. Subsequently, lysates were homogenized by sonication (6 × 10 sec, 15 W),
cleared by centrifugation (20,000 × g, 15 min, 4 °C), and the supernatant pre-
cipitated by addition of four volumes of 20 °C acetone. Precip itates were re-
dissolved in 8 M guanidine HCl, 50 mM Hepes pH8.5 and protein concentration
was determined by Quick-start Bradford assay (Bio-Rad).
Chemical acetylation. Protein lysates in 8 M guanidine HCl were mixed with 1/10
volume 1 M acetyl-phosphate (Sigma) prepared in H
2
O, and the acetylation
reaction was allowed to proceed for 2 h at 37 °C. Control reactions were prepared
by mixing with 1/10 volume H
2
O. The acetyl-phosphate was quenched by diluting
the reaction ve-fold in 8 M Guanidine HCl, 100 mM Tris pH8. Control and
chemically acetylated SILAC-labeled protein lysates were mixed in equal portions
according to the protein concentration as determined by Bradford assay (above).
The samples were digested (described below) and analyzed by mass spectrometry
to determine the degree of partial chemical acetylation (PCA) (Supplementary
Figure 1a). The measured SILAC ratios were additionally used to adjust the mixing
of SILAC lysates in subsequent experiments.
Protein digestion. Protein lysates were reduced and alkalated with 5 mM TCEP
and 5 mM chloracetamide for 45 min at room temperature. Approximately 10 mg
protein per condition was diluted to 2 M guanidine HCl with 50 mM Hepes, pH8
and digested by endoproteinase Lys C (1:200 w/w; Wako) for 2 h at room tem-
perature. The lysates were further diluted to 1 M guanidine HCl and digested with
trypsin protease (1:200 w/w; Sigma-Aldrich) for 16 h at 37 °C. Digestion was
stopped by the addition of triuoroacetic acid (TFA) to a nal concentration of 1%.
Digests were cleared by centrifugation (2500 × g, 5 min) and loaded onto reversed-
phase C18 Sep-Pak columns (Waters), pre-equilibrated with 5 ml acetonitrile and
2 × 5 ml 0.1% TFA. Peptides were washed with 0.1% TFA and H
2
O, and eluted
with 50% acetonitrile.
Acetylated peptide enrichment. Peptide concentration was determined by UV
spectrometry. The rst experiment was performed in technical replicates using two
different peptide fractionation strategies. In the rst strategy, peptides were pre-
fractionated by high pH reversed-phase chromatography11 to 6 fractions followed
by acetylated peptide enrichment and micro-scale (in stage-tip) scale strong cation
exchange (SCX) chromatography to an additional 3 fractions (pH4.5, 5.5, and 9.0),
for a total of 18 fractions. The second strategy was to perform acetylated peptide
enrichment on the entire peptide digest, followed by micro-scale SCX into 5
fractions (pH3.5, 4.0, 4.5, 5.5, and 9.0). The second experiment was performed
using the second fractionation strategy only. For acetylated peptide enrichment, the
peptides were mixed with 100 µl 10x IP buffer (500 mM MOPS; pH 7.2, 100 mM
Na-phosphate, 500 mM NaCl, 5% NP-40) per 5 mg peptides. The acetonitrile was
removed and the volume reduced to ~1 ml, by vacuum centrifugation. The nal
volume was adjusted with H
2
O to a concentration of 5 mg/ml. 40 μl of anti-
acetylated lysine antibody (PTMScan Acetyl-Lysine Motif [Ac-K] Kit, Cell Sig-
naling Technology) was washed in 1 mL IP buffer, the peptides were claried by
centrifugation at 20,000 × gfor 5 min, and the peptide supernatant was mixed with
the anti-acetylated lysine antibody. Peptides were enriched overnight at 4 °C,
washed 3 × in 1 ml cold (4 °C) IP buffer, in 1 ml cold IP buffer without NP-40,
and in 1 ml H
2
O. All wash buffer was removed using a 26 gauge needle on an
aspirator. Acetylated peptides were eluted with 100 μl 0.15% TFA, repeated for a
total of three times. Peptides were loaded directly onto a micro-SCX column,
fractionated as described above, and de-salted on C18 stage-tips38.
Mass spectrometry. Peptides were analyzed by nanoow liquid chromatography-
coupled tandem mass spectrometry (nLC-MS/MS) using a Proxeon easy nLC 1200
connected to a Q-Exactive HF mass spectrometer (Thermo Scientic). The Q-
Exactive was operated in prole mode using positive polarity and a Top10 data
dependent acquisition (DDA) method with the following settings: Spray voltage =
2 kV, S-lens RF level =50, heated capillary =275 °C. Full scan (MS1) was per-
formed with an m/z range of 3001750 at 60,000 resolution with a target value of
3×10
6ions and a maximum ll time of 20 milliseconds (ms). Fragment (MS2)
scans were performed at a resolution of 15,000 with a target value of 5 × 104ions, a
maximum injection time of 110 ms, an isolation width of 1.3 m/z, a normalized
collision energy (NCE) of 28, and a xed rst mass of 100 m/z. Peptide were
separated by nanoow chromatography using an EASY-nLC 1000 system (Thermo
Scientic) connected to a 15 cm capillary column packed with 1.9 μm Reprosil-Pur
C18 beads (Dr. Maisch). Column temperat ure was maintained at 40 °C using an
integrated column oven (PRSO-V1, Sonation GmbH) Peptides were eluted by a
gradient of acetonitrile (ACN) in 0.1% formic acid. A typical run utilized a 120 min
gradient followed by 15 min wash and equilibration. A linear gradient at 250 nl/
min of 8-24% ACN (90 min) and 2440% ACN (30 min) was followed by a wash at
4080% ACN (5 min) and 80%8% ACN (5 min), and equilibration at 8% ACN (5
min).
Raw MS data were analyzed using MaxQuant (developer version 1.5.5.4) with
the integrated Andromeda search engine39 to search the UniProt human FASTA
(downloaded 6 July 2015). The following Andromeda settings were used; initial
search mass tolerance of 20 ppm, main search mass tolerance of 6 ppm for parent
ions and 20 ppm for HCD fragment ions, trypsin specicity with a maximum of
two missed cleavages, cysteine carbamidomethylation as a xed modication, and
N-acetylation of proteins, oxidized methionine, and acetylated lysine as variable
modications. Acetylated peptides were ltered for a minimum Andromeda score
of 40, as per the default settings for modied peptides. Known contaminants were
removed based on classication by MaxQuant. The false discovery rate (FDR) was
estimated for peptides and proteins individually using a target-decoy approach
allowing a maximum of 1% false identications from a reversed sequence database.
Calculation of acetylation stoichiometry. We used MaxQuant to analyze the
SILAC ratios of native and chemically acetylated peptides. In order to accurately
calculate stoichiometry it is important to compare the SILAC ratios of individual,
singly acetylated peptides from the evidence.txttable. We do not use the Acetyl
(K)Sites.txttable since the SILAC ratios of individual sites are sometimes derived
from multiple peptides. Likewise, entries in the modicationSpecicPeptides.txt
table can include different positions of acetylation within the same peptide
sequence. Starting with the evidence.txttable the following actions were
performed;
1. Remove all entries where Reverse =+, Potential contaminant =+, Acetyl
(K) =0, and Acetyl (K) =2. This removes reverse and contaminant entries
and results in only entries containing singly acetylated peptides.
2. The Modied Sequence was used as a unique identier to calculate the median
SILAC ratio and summed peptide intensity for multiple instances of any given
acetylated peptide in each experiment.
3. SILAC ratios were tested for agreement with the dilution series, allowing for
up to two-fold variability8. Stoichiometry was calculated using only peptides
that met these strict criteria.
4. Stoichiometry was calculated as follows; Stoichiometry (S), degree partial
chemical acetylation (C), dilution factor for acetylated peptides (D), and
SILAC ratio partial chemical acetylation/native acetylation (R). S =(C)/
((R*D)-(1-C)). The dilution factors (D) were as follows: Experiment 1, (~1%
D=4.23), (~0.1% D =42.3), (~0.01% D =423), Experiment 2, (~1% D =
6.37), (~0.1% D =63.7), (~0.01% D =637). The median degree of partial
chemical acetylation (C) was 3.53% and 10.38%, for experiments 1 and 2,
respectively, (Supplementary Figure 1a).
The above mentioned .txt les can be found via the PRIDE40 partner repository
with the dataset identier PXD009994. Relative acetylation stoichiometry was
estimated using acetylation site intensity corrected for differences in protein
abundance, otherwise referred to as abundance-corrected intensity (ACI). ACI was
calculated by dividing acetylation site intensity by iBAQ protein abundance.
Validation by AQUA and spike-in. Two different unmodied AQUA peptides
(Thermo Scientic, AQUA quant pro) per protein were added to SILAC heavy-
labeled HeLa peptides to determine the concentration of each protein (CTTN,
NCL, and NAT10). Acetylated AQUA peptides were then added to nal stoi-
chiometry of 1%, 0.1%, or 0.01%, in accordance with each individual proteins
concentration. Acetylated peptides were enriched as described above and analyzed
by mass spectrometry. For validation by recombinant acetylated protein spike-in,
recombinant acetylated protein was added to HeLa lysate and digested to deter-
mine the concentration of protein present in the lysate. Based on the initial ana-
lysis, recombinant HMGCL K48ac was added at a 1:1 (100%), 1:10 (10%), and
1:100 (1%) stoichiometry and MDH2 K239ac was added at a 1:1 (100%), 1:1,000
(0.1%), and 1:10,000 (0.01%) stoichiometry. Acetylated peptides were enriched
from the 1:10 and 1:100 dilutions (HMGCL) and the 1:1,000 and 1:10,000 dilutions
(MDH2) before analysis by mass spectrometry. The 1:1 dilutions were analyzed to
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-09024-0 ARTICLE
NATURE COMMUNICATIONS | (2019) 10:1055 | https://doi.org/10.1038/s41467-019-09024-0 | www.nature.com/naturecommunications 9
Content courtesy of Springer Nature, terms of use apply. Rights reserved
determine the protein mixing ratio and to correct the stoichiometry measurements
accordingly. The mixing ratio of spike-in/HeLa was 1.48 for HMGCL and 0.70 for
MDH2.
Expression and purication of recombinant acetylated proteins. MDH2
(K239AcK) and HMGCL (K48AcK) were expressed from a pRSF-Duet-vector
containing the HMGCL coding region with an amber-stop codon at the position
encoding for acetyl-L-lysine incorporation, and additionally the coding regions for
a synthetically evolved pair of Methanosarcina barkeri MS tRNA
CUA
(MbtRNA-
CUA) and the acetyl-lysyl-tRNA-sythetase (AcKRS3). AcKRS3 and MbtRNA
CUA
enable the site-specic incorporation of acetyl-L-lysine into proteins as response to
an amber codon. This was done by supplementing the E. coli BL21 (DE3) culture
with 10 mM N-(e)-acetyl-lysine (Chem-Impex International, Inc.) and 20 mM
nicotinamide to inhibit the E. coli CobB deacetylase at an OD
600
of 0.6 (37 °C, 160
rpm). Subsequently, the temperat ure was lowered to 18 °C and cells were grown for
another 30 min at 160 rpm. Protein expression was induced by addition of 200 µM
IPTG and protein expression conducted overnight (18 °C, 160 rpm). The cells were
harvested by centrifugation (4000 × g, 20 min) and resuspended in buffer A. Cell
lysis and protein purication was performed as described above.
Histone western blot. HeLa cells were lysed by boiling in 2% lithium dodecyl
sulfate (LDS) (1x NuPAGE LDS Sample Buffer, ThermoFischer Scientic), fol-
lowed by sonication to disrupt genomic DNA. Proteins were separated on a 4-12
NuPAGE gel (ThermoFischer Scientic) and transferred to a nitrocellulose
membrane (BioRad). Acetylated proteins were visualized by immunoblot using
pan-anti-acetylated lysine antibody (Immunechem #ICP0380) at a 1/1000 dilution
and goat anti-rabbit horseradisch peroxidase (HRP) (BioRad, STAR124P) sec-
ondary antibody at a 1/5000 dilution.
Data analysis. Pearsons correlation (r) and Wilcoxon tests were performed in R.
UniProt keyword analysis was performed using AGOTOOL10 with an uncorrected
P-value cutoff of 0.05. Word clouds were generated in R. To estimate histone
acetylation copy numbers we used the copy number estimate for histone H4 (5e7
per cell) for all four core histones. This is because copy number estimates for
histones are complicated by the presence of multiple isoforms of H2B, H2A, and
H3. The copy number estimate for H4 is consistent with a previous analysis that
estimated that core histones represent ~4% of the total protein in HeLa cells, and
with the DNA content of HeLa cells41. We furthermore used the median stoi-
chiometry for sites that occurred on multiple histone isoforms (such as H2B K5) to
avoid over-counting these sites.
Reporting summary. Further information on experimental design is available in
the Nature Research Reporting Summary linked to this article.
Data availability
The raw mass spectrometry data have been deposited to the ProteomeXchange
Consortium via the PRIDE40 partner repository with the dataset identier PXD009994.
The source data underlying Figs. 1a, 2a-b, 2d-e, 3a-d, 4a, 4c-d, 5a-c and Supplementary
Figs. 1a-c, 2a-c, and 3 are provided as a Source Data le. A reporting summary for this
Article is available as a Supplementary Information le. All other data supporting the
ndings of this study are available from the corresponding authors upon reasonable
request.
Received: 20 November 2018 Accepted: 11 February 2019
References
1. Allfrey, V. G., Faulkner, R. & Mirsky, A. E. Acetylation and methylation of
histones and their possible role in the regulation of RNA synthesis. Proc. Natl
Acad. Sci. USA 51, 786794 (1964).
2. Kim, S. C. et al. Substrate and functional diversity of lysine acetylation
revealed by a proteomics survey. Mol. Cell 23, 607618 (2006).
3. Choudhary, C. et al. Lysine acetylation targets protein complexes and co-
regulates major cellular functions. Science 325, 834840 (2009).
4. Lundby, A. et al. Proteomic analysis of lysine acetylation sites in rat tissues
reveals organ specicity and subcellular patterns. Cell Rep. 2, 419431 (2012).
5. Wagner, G. R., Payne, R. M. Widespread & Enzyme-independent, N {epsilon}-
Acetylation and N{epsilon}-succinylation of proteins in the chemical
conditions of the mitochondrial matrix. J. Biol. Chem. 288, 2903629045
(2013).
6. Weinert, B. T. et al. Acetylation dynamics and stoichiometry in
Saccharomyces cerevisiae. Mol. Syst. Biol. 10, 716 (2014).
7. Weinert, B. T., Moustafa, T., Iesmantavicius, V., Zechner, R. & Choudhary, C.
Analysis of acetylation stoichiometry suggests that SIRT3 repairs
nonenzymatic acetylation lesions. EMBO J. 34, 26202632 (2015).
8. Weinert, B. T. et al. Accurate quantication of site-specic acetylation
stoichiometry reveals the impact of sirtuin deacetylase CobB on the E. coli
acetylome. Mol. Cell. Proteom. 16, 759769 (2017).
9. Carabetta V. J., Greco T. M., Tanner A. W., Cristea I. M., Dubnau D.
Temporal regulation of the Bacillus subtilis acetylome and evidence for a role
of MreB acetylation in cell wall growth. mSystems 1, e0000516 (2016).
10. Scholz, C. et al. Avoiding abundance bias in the functional annotation of
posttranslationally modied proteins. Nat. Methods 12, 10031004 (2015).
11. Svinkina, T. et al. Deep, quantitative coverage of the lysine acetylome using
novel anti-acetyl-lysine antibodies and an optimized proteomic workow.
Mol. Cell. Proteom. 14, 24292440 (2015).
12. Weinert, B. T. et al. Time-resolved analysis reveals rapid dynamics and broad
scope of the CBP/p300 acetylome. Cell 174, 231244 e212 (2018).
13. Bekker-Jensen, D. B. et al. An optimized shotgun strategy for the rapid
generation of comprehensive human proteomes. Cell Syst 4, 587599 (2017).
14. Thul P. J., et al. A subcellular map of the human proteome. Science
356, eaal3321 (2017).
15. Colaert, N., Helsens, K., Martens, L., Vandekerckhove, J. & Gevaert, K.
Improved visualization of protein consensus sequences by iceLogo. Nat.
Methods 6, 786787 (2009).
16. James, A. M., Smith, A. C., Smith, C. L., Robinson, A. J. & Murphy, M. P.
Proximal cysteines that enhance lysine N-acetylation of cytosolic proteins in
mice are less conserved in longer-living species. Cell Rep. 24, 14451455
(2018).
17. Abshiru, N. et al. Discovery of protein acetylation patterns by deconvolution
of peptide isomer mass spectra. Nat. Commun. 6, 8648 (2015).
18. Feller, C., Forne, I., Imhof, A. & Becker, P. B. Global and specic responses of
the histone acetylome to systematic perturbation. Mol. Cell 57, 559571
(2015).
19. Zhou, T., Chung, Y. H., Chen, J. & Chen, Y. Site-specic identication of
lysine acetylation stoichiometries in mammalian cells. J. Proteome Res. 15,
11031113 (2016).
20. Zheng, Y., Thomas, P. M. & Kelleher, N. L. Measurement of acetylation
turnover at distinct lysines in human histones identies long-lived acetylation
sites. Nat. Commun. 4, 2203 (2013).
21. Scholz, C. et al. Acetylation site specicities of lysine deacetylase inhibitors in
human cells. Nat. Biotechnol. 33, 415423 (2015).
22. Dancy, B. M. & Cole, P. A. Protein lysine acetylation by p300/CBP. Chem. Rev.
115, 24192452 (2015).
23. Ishfaq, M. et al. Acetylation regulates subcellular localization of eukaryotic
translation initiation factor 5A (eIF5A). FEBS Lett. 586, 32363241 (2012).
24. Du, Z. et al. DNMT1 stability is regulated by proteins coordinating
deubiquitination and acetylation-driven ubiquitination. Sci. Signal. 3, ra80
(2010).
25. Fan, J. et al. Tyr phosphorylation of PDP1 toggles recruitment between
ACAT1 and SIRT3 to regulate the pyruvate dehydrogenase complex. Mol. Cell
53, 534548 (2014).
26. Wang, Y. P. et al. Regulation of G6PD acetylation by SIRT2 and KAT9
modulates NADPH homeostasis and cell survival during oxidative stress.
EMBO J. 33, 13041320 (2014).
27. Wang, S. et al. Insulin and mTOR pathway regulate HDAC3-mediated
deacetylation and activation of PGK1. PLoS Biol. 13, e1002243 (2015).
28. Mateo, F. et al. The transcriptional co-activator PCAF regulates cdk2 activity.
Nucleic Acids Res. 37, 70727084 (2009).
29. Xia, P. et al. EB1 acetylation by P300/CBP-associated factor (PCAF) ensures
accurate kinetochore-microtubule interactions in mitosis. Proc. Natl Acad. Sci.
USA 109, 1656416569 (2012).
30. Lin, H., Su, X. & He, B. Protein lysine acylation and cysteine succination
by intermediates of energy metabolism. Acs. Chem. Biol. 7, 947960
(2012).
31. Wagner, G. R. & Hirschey, M. D. Nonenzymatic protein acylation as a carbon
stress regulated by sirtuin deacylases. Mol. Cell 54,516 (2014).
32. Tatham, M. H. et al. A Proteomic approach to analyze the aspirin-mediated
lysine acetylome. Mol. Cell. Proteom. 16, 310326 (2017).
33. Baeza, J. et al. Stoichiometry of site-specic lysine acetylation in an entire
proteome. J. Biol. Chem. 289, 2132621338 (2014).
34. Gil, J. et al. Lysine acetylation stoichiometry and proteomics analyses reveal
pathways regulated by sirtuin 1 in human cells. J. Biol. Chem. 292,
1812918144 (2017).
35. Meyer, J. G. et al. Quantication of lysine acetylation and succinylation
stoichiometry in proteins using mass spectrometric data-independent
acquisitions (SWATH). J. Am. Soc. Mass. Spectrom. 27, 17581771 (2016).
36. Bakalarski, C. E. et al. The impact of peptide abundance and dynamic range
on stable-isotope-based quantitative proteomic analyses. J. Proteome Res. 7,
47564765 (2008).
37. Lau, H. T., Suh, H. W., Golkowski, M. & Ong, S. E. Comparing SILAC- and
stable isotope dimethyl-labeling approaches for quantitative proteomics. J.
Proteome Res. 13, 41644174 (2014).
ARTICLE NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-09024-0
10 NATURE COMMUNICATIONS | (2019) 10:1055 | https://doi.org/10.1038/s41467-019-09024-0 | www.nature.com/naturecommunications
Content courtesy of Springer Nature, terms of use apply. Rights reserved
38. Rappsilber, J., Ishihama, Y. & Mann, M. Stop and go extraction tips for
matrix-assisted laser desorption/ionization, nanoelectrospray, and LC/MS
sample pretreatment in proteomics. Anal. Chem. 75, 663670 (2003).
39. Cox, J. et al. Andromeda: a peptide search engine integrated into the
MaxQuant environment. J. Proteome Res. 10, 17941805 (2011).
40. Vizcaino, J. A. et al. The PRoteomics IDEntications (PRIDE) database and
associated tools: status in 2013. Nucleic Acids Res. 41, D1063D1069 (2013).
41. Wisniewski, J. R., Hein, M. Y., Cox, J. & Mann, M. A. proteomic rulerfor
protein copy number and concentration estimation without spike-in
standards. Mol. Cell. Proteom. 13, 34973506 (2014).
42. Kulak, N. A., Pichler, G., Paron, I., Nagaraj, N. & Mann, M. Minimal,
encapsulated proteomic-sample processing applied to copy-number
estimation in eukaryotic cells. Nat. Methods 11, 319324 (2014).
Acknowledgements
We thank the members of our laboratory for their helpful discussions. L.B. and M.L. were
funded by the German Research Foundation (CECAD and DFG Research Grant Number
LA2984/5-1). C.C. is supported by the Hallas Møller Investigator award (NNF14OC0008541)
from the Novo Nordisk. B.T.W. is supported by a grant from the Novo Nordisk Foundation
(NNF15OC0017774). The Novo Nordisk Foundation Center for Protein Research is sup-
ported nancially by the Novo Nordisk Foundation (Grant agreement NNF14CC0001).
Author contributions
B.T.W. and C.C. designed the research and supervised the project. B.K.H. performed
stoichiometry measurements and contributed to data interpretation. R.G. performed
validation experiments. L.B and M.L. provided recombinant acetylated proteins. D.L. and
T.N. provided bioinformatics support. B.T.W. analyzed data, prepared gures, and wrote
the manuscript. All authors read and commented on the manuscript.
Additional information
Supplementary Information accompanies this paper at https://doi.org/10.1038/s41467-
019-09024-0.
Competing interests: The authors declare no competing interests.
Reprints and permission information is available online at http://npg.nature.com/
reprintsandpermissions/
Journal peer review information:Nature Communications thanks Xiao-Lu Zhao and the
other anonymous reviewers for their contribution to the peer review of this work. Peer
reviewer reports are available.
Publishers note: Springer Nature remains neutral with regard to jurisdictional claims in
published maps and institutional afliations.
Open Access This article is licensed under a Creative Commons
Attribution 4.0 International License, which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give
appropriate credit to the original author(s) and the source, provide a link to the Creative
Commons license, and indicate if changes were made. The images or other third party
material in this article are included in the articles Creative Commons license, unless
indicated otherwise in a credit line to the material. If material is not included in the
articles Creative Commons license and your intended use is not permitted by statutory
regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder. To view a copy of this license, visit http://creativecommons.org/
licenses/by/4.0/.
© The Author(s) 2019
NATURE COMMUNICATIONS | https://doi.org/10.1038/s41467-019-09024-0 ARTICLE
NATURE COMMUNICATIONS | (2019) 10:1055 | https://doi.org/10.1038/s41467-019-09024-0 | www.nature.com/naturecommunications 11
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... Acetylation of the ε-amino group of the lysine side chain is a post-translational modification that occurs on thousands of proteins. However, the bulk (74%) of this acetylation is found on histone proteins with their lysine-rich tails (1,2). Lysine acetylation on the N-terminal tails of histone proteins is an essential mechanism for the regulation of chromatin structure and gene expression (3). ...
... In contrast, double knockout (DKO) of Hdac1/2 produces severe phenotypes, which is unsurprising given that HDAC1/2 are responsible for over half of the total deacetylase activity in cells (9,21). The use of both genetic knockouts (KOs) and class I HDAC specific inhibitors has highlighted the importance of HDAC1/2 in removing acetylation from high stoichiometry sites on histone tails (1,9,21). Despite the removal of acetyl marks appearing to be a repressive function, ChIP-seq revealed that HDAC1/2 and their associated complexes are found located at sites of active transcription (22,23). ...
... As 6 of these 7 corepressor complexes contain HDAC1 (and/or HDAC2) it is no surprise that degradation of HDAC1 revealed a wide range of substrates across the core histones (Fig 2B). Many of the sites identified are high stoichiometry acetylation sites, including H4K5, H4K8, H4K12, H4K16, H2BK5 and H2BK20 (1,(53)(54)(55)(56). Strikingly, several of the sites (including those on H2B) are targets for the crucial acetyltransferases, p300/CBP (39). ...
Preprint
Full-text available
Histone Deacetylase 1 (HDAC1) removes acetyl groups from lysine residues on the core histones, a critical step in the regulation of chromatin accessibility. Despite histone deacetylation being an apparently repressive activity, suppression of HDACs causes both up- and down-regulation of gene expression. Here we exploited the degradation tag (dTAG) system to rapidly degrade HDAC1 in embryonic stem cells (ESCs) lacking its paralog, HDAC2. Unlike HDAC inhibitors that lack isoform specificity, the dTAG system allowed specific degradation and removal of HDAC1 in <1 hour (100x faster than genetic knockouts). This rapid degradation caused increased histone acetylation in as little as 2 hours, with H2BK5 and H2BK11 being the most sensitive. The majority of differentially expressed genes following 2 hours of HDAC1 degradation were upregulated (275 genes up vs 15 down) with increased proportions of downregulated genes observed at 6 (1,153 up vs 443 down) and 24 hours (1,146 up vs 967 down) respectively. Upregulated genes showed increased H2BK5ac and H3K27ac around their transcriptional start site (TSS). In contrast, decreased acetylation of super-enhancers (SEs) was linked to the most strongly downregulated genes. These findings suggest a paradoxical role for HDAC1 in the maintenance of histone acetylation levels at critical enhancer regions required for the pluripotency-associated gene network.
... Among acylation modifications, lysine acetylation is the simplest and best characterized type. It has been reported in both yeast and mammals that most acetylation lysine sites are of low stoichiometries (<1%) [14][15][16][17][18] . These observations question the functional importance of most acetylation sites and cast doubts on the biological relevance of lysine acylations in general. ...
... Meanwhile, by using partial chemical acetylation and stable isotope labelling by amino acids (SILAC)-based quantification, Weinert et al. reported that 95% of acetylation sites had a stoichiometry <1% in exponentially growing yeast 14 . With the same approach, they also estimated that the median stoichiometry of acetylation was only ~0.05% in mouse liver 15 , and ~0.02% in HeLa cells 16 . These studies suggest that the majority of acetylated lysine sites are of very low acetylation stoichiometries, and not biologically important. ...
Article
Full-text available
The post-translational modification lysine succinylation is implicated in the regulation of various metabolic pathways. However, its biological relevance remains uncertain due to methodological difficulties in determining high-impact succinylation sites. Here, using stable isotope labelling and data-independent acquisition mass spectrometry, we quantified lysine succinylation stoichiometries in mouse livers. Despite the low overall stoichiometry of lysine succinylation, several high-stoichiometry sites were identified, especially upon deletion of the desuccinylase SIRT5. In particular, multiple high-stoichiometry lysine sites identified in argininosuccinate synthase (ASS1), a key enzyme in the urea cycle, are regulated by SIRT5. Mutation of the high-stoichiometry lysine in ASS1 to succinyl-mimetic glutamic acid significantly decreased its enzymatic activity. Metabolomics profiling confirms that SIRT5 deficiency decreases urea cycle activity in liver. Importantly, SIRT5 deficiency compromises ammonia tolerance, which can be reversed by the overexpression of wild-type, but not succinyl-mimetic, ASS1. Therefore, lysine succinylation is functionally important in ammonia metabolism.
... The ~1% stoichiometry level of K22 and K95-CaM acetylation in the total lysates of forebrain was similar to those of bona fide acetylated proteins in mammalian cells. 58 The calcium elevation during learning is localized to stimulated spines, or a region called calcium nanodomain which is near the inner mouth of postsynaptic NMDA receptor. 28 One could speculate that the stoichiometry level of CaM acetylation in the stimulated spines or calcium nanodomain might be much higher than that in the total lysates. ...
Article
Full-text available
Aims Alzheimer's disease (AD) is a neurodegenerative disease characterized by progressive cognitive dysfunction and memory impairment. AD pathology involves protein acetylation. Previous studies have mainly focused on histone acetylation in AD, however, the roles of nonhistone acetylation in AD are less explored. Methods The protein acetylation and expression levels were detected by western blotting and co‐immunoprecipitation. The stoichiometry of acetylation was measured by home‐made and site‐specific antibodies against acetylated‐CaM (Ac‐CaM) at K22, K95, and K116. Hippocampus‐dependent learning and memory were evaluated by using the Morris water maze, novel object recognition, and contextual fear conditioning tests. Results We showed that calmodulin (CaM) acetylation is reduced in plasma of AD patients and mice. CaM acetylation and its target Ca²⁺/CaM‐dependent kinase II α (CaMKIIα) activity were severely impaired in AD mouse brain. The stoichiometry showed that Ac‐K22, K95‐CaM acetylation were decreased in AD patients and mice. Moreover, we screened and identified that lysine deacetylase 9 (HDAC9) was the main deacetylase for CaM. In addition, HDAC9 inhibition increased CaM acetylation and CaMKIIα activity, and hippocampus‐dependent memory in AD mice. Conclusions HDAC9‐mediated CaM deacetylation induces memory impairment in AD, HDAC9, or CaM acetylation may become potential therapeutic targets for AD.
... full-length human STAT1 protein (hSTAT1) with acetic anhydride and incubate it with human recombinant HDAC6 (hHDAC6) after purification to detect changes in the degree of acetylation (Fig. S3) [35,36]. With the increase of acetic anhydride concentration, the acetylation level of STAT1 protein gradually increased and was relatively highest when the acetic anhydride concentration was 6 mM (Fig. S3A). ...
Article
Full-text available
The search for effective combination therapy with immune checkpoint inhibitors (ICI) has become important for cancer patients who do not respond to the ICI well. Histone deacetylases (HDACs) inhibitors have attracted wide attention as anti-tumor agents. ACY-1215 is a selective inhibitor of HDAC6, which can inhibit the growth of a variety of tumor. We previously revealed that HDAC family is highly expressed in colorectal cancer specimens and mouse models. In this study, ACY-1215 was combined with anti-PD1 to treat tumor-bearing mice associated with colorectal cancer. ACY-1215 combined with anti-PD1 effectively inhibited the colorectal tumor growth. The expression of PD-L1 in tumor of mice were inhibited by ACY-1215 and anti-PD1 combination treatment, whereas some biomarkers reflecting T cell activation were upregulated. In a co-culture system of T cells and tumor cells, ACY-1215 helped T cells to kill tumor cells. Mechanically, HDAC6 enhanced the acetylation of STAT1 and inhibited the phosphorylation of STAT1, thus preventing STAT1 from entering the nucleus to activate PD-L1 transcription. This study reveals a novel regulatory mechanism of HDAC6 on non-histone substrates, especially on protein acetylation. HDAC6 inhibitors may be of great significance in tumor immunotherapy and related combination strategies. Supplementary Information The online version contains supplementary material available at 10.1007/s00262-023-03624-y.
Article
Full-text available
In all eukaryotes, acetylation of histone lysine residues correlates with transcription activation. Whether histone acetylation is a cause or consequence of transcription is debated. One model suggests that transcription promotes the recruitment and/or activation of acetyltransferases, and histone acetylation occurs as a consequence of ongoing transcription. However, the extent to which transcription shapes the global protein acetylation landscapes is not known. Here, we show that global protein acetylation remains virtually unaltered after acute transcription inhibition. Transcription inhibition ablates the co-transcriptionally occurring ubiquitylation of H2BK120 but does not reduce histone acetylation. The combined inhibition of transcription and CBP/p300 further demonstrates that acetyltransferases remain active and continue to acetylate histones independently of transcription. Together, these results show that histone acetylation is not a mere consequence of transcription; acetyltransferase recruitment and activation are uncoupled from the act of transcription, and histone and non-histone protein acetylation are sustained in the absence of ongoing transcription.
Article
Ubiquitylation regulates most proteins and biological processes in a eukaryotic cell. However, the site-specific occupancy (stoichiometry) and turnover rate of ubiquitylation have not been quantified. Here we present an integrated picture of the global ubiquitylation site occupancy and half-life. Ubiquitylation site occupancy spans over four orders of magnitude, but the median ubiquitylation site occupancy is three orders of magnitude lower than that of phosphorylation. The occupancy, turnover rate, and regulation of sites by proteasome inhibitors are strongly interrelated, and these attributes distinguish sites involved in proteasomal degradation and cellular signaling. Sites in structured protein regions exhibit longer half-lives and stronger upregulation by proteasome inhibitors than sites in unstructured regions. Importantly, we discovered a surveillance mechanism that rapidly and site-indiscriminately deubiquitylates all ubiquitin-specific E1 and E2 enzymes, protecting them against accumulation of bystander ubiquitylation. The work provides a systems-scale, quantitative view of ubiquitylation properties and reveals general principles of ubiquitylation-dependent governance.
Article
Full-text available
The interplay between genetic alterations and metabolic dysregulation is increasingly recognized as a pivotal axis in cancer pathogenesis. Both elements are mutually reinforcing, thereby expediting the ontogeny and progression of malignant neoplasms. Intriguingly, recent findings have highlighted the translocation of metabolites and metabolic enzymes from the cytoplasm into the nuclear compartment, where they appear to be intimately associated with tumor cell proliferation. Despite these advancements, significant gaps persist in our understanding of their specific roles within the nuclear milieu, their modulatory effects on gene transcription and cellular proliferation, and the intricacies of their coordination with the genomic landscape. In this comprehensive review, we endeavor to elucidate the regulatory landscape of metabolic signaling within the nuclear domain, namely nuclear metabolic signaling involving metabolites and metabolic enzymes. We explore the roles and molecular mechanisms through which metabolic flux and enzymatic activity impact critical nuclear processes, including epigenetic modulation, DNA damage repair, and gene expression regulation. In conclusion, we underscore the paramount significance of nuclear metabolic signaling in cancer biology and enumerate potential therapeutic targets, associated pharmacological interventions, and implications for clinical applications. Importantly, these emergent findings not only augment our conceptual understanding of tumoral metabolism but also herald the potential for innovative therapeutic paradigms targeting the metabolism–genome transcriptional axis.
Article
Full-text available
Acetyl-coenzyme A (CoA) is an abundant metabolite that can also alter protein function through non-enzymatic N-acetylation of protein lysines. This N-acetylation is greatly enhanced in vitro if an adjacent cysteine undergoes initial S-acetylation, as this can lead to S→N transfer of the acetyl moiety. Here, using modeled mouse structures of 619 proteins N-acetylated in mouse liver, we show lysine N-acetylation is greater in vivo if a cysteine is within ∼10 Å. Extension to the genomes of 52 other mammalian and bird species shows pairs of proximal cysteine and N-acetylated lysines are less conserved, implying most N-acetylation is detrimental. Supporting this, there is less conservation of cytosolic pairs of proximal cysteine and N-acetylated lysines in species with longer lifespans. As acetyl-CoA levels are linked to nutrient supply, these findings suggest how dietary restriction could extend lifespan and how pathologies resulting from dietary excess may occur.
Article
Full-text available
Lysine acetylation is a widespread posttranslational modification (PTM) affecting many biological pathways. Recent studies indicate that acetylated lysine residues mainly exhibit low acetylation occupancy, but challenges in sample preparation and analysis make it difficult to confidently assign these numbers, limiting understanding of their biological significance. Here, we tested three common sample preparation methods to determine their suitability for assessing acetylation stoichiometry in three human cell lines, identifying the acetylation occupancy in more than 1,300 proteins from each cell line. The stoichiometric analysis in combination with quantitative proteomics also enabled us to explore their functional roles. We found that higher abundance of the deacetylase sirtuin 1 (SIRT1) correlated with lower acetylation occupancy and lower levels of ribosomal proteins including those involved in ribosome biogenesis and rRNA processing. Treatment with the SIRT1 inhibitor EX-527 confirmed SIRT1’s role in the regulation of pre-rRNA synthesis and processing. Specifically, proteins involved in pre-rRNA transcription, including subunits of the Pol 1 and SL1 complexes and the RNA polymerase I specific transcription initiation factor RRN3 were up-regulated after SIRT1 inhibition. Moreover, many protein effectors and regulators of pre-rRNA processing needed for rRNA maturation were also up-regulated after EX-527 treatment, with the outcome that pre-rRNA and 28S rRNA levels also increased. More generally, we found that SIRT1 inhibition down-regulates metabolic pathways including glycolysis and pyruvate metabolism. Together, these results provide the largest dataset thus far of lysine acetylation stoichiometry (available via ProteomeXchange with identifier PXD005903) and set the stage for further biological investigations of this central PTM.
Article
Full-text available
This study investigates the challenge of comprehensively cataloging the complete human proteome from a single-cell type using mass spectrometry (MS)-based shotgun proteomics. We modify a classical two-dimensional high-resolution reversed-phase peptide fractionation scheme and optimize a protocol that provides sufficient peak capacity to saturate the sequencing speed of modern MS instruments. This strategy enables the deepest proteome of a human single-cell type to date, with the HeLa proteome sequenced to a depth of ∼584,000 unique peptide sequences and ∼14,200 protein isoforms (∼12,200 protein-coding genes). This depth is comparable with next-generation RNA sequencing and enables the identification of post-translational modifications, including ∼7,000 N-acetylation sites and ∼10,000 phosphorylation sites, without the need for enrichment. We further demonstrate the general applicability and clinical potential of this proteomics strategy by comprehensively quantifying global proteome expression in several different human cancer cell lines and patient tissue samples.
Article
Full-text available
Lysine acetylation is a protein posttranslational modification (PTM) that occurs on thousands of lysine residues in diverse organisms from bacteria to humans. Accurate measurement of acetylation stoichiometry on a proteome-wide scale remains challenging. Most methods employ a comparison of chemically acetylated peptides to native acetylated peptides, however, the potentially large differences in abundance between these peptides presents a challenge for accurate quantification. Stable isotope labeling by amino acids in cell culture (SILAC)-based mass spectrometry (MS) is one of the most widely used quantitative proteomic methods. Here we show that serial dilution of SILAC-labeled peptides (SD-SILAC) can be used to identify accurately quantified peptides and to estimate the quantification error rate. We applied SD-SILAC to determine absolute acetylation stoichiometry in exponentially-growing and stationary-phase wild type and Sirtuin deacetylase CobB-deficient cells. To further analyze CobB-regulated sites under conditions of globally increased or decreased acetylation, we measured stoichiometry in phophotransacetylase (ptaΔ) and acetate kinase (ackAΔ) mutant strains in the presence and absence of the Sirtuin inhibitor nicotinamide. We measured acetylation stoichiometry at 3,669 unique sites and found that the vast majority of acetylation occurred at a low stoichiometry. Manipulations that cause increased nonenzymatic acetylation by acetyl-phosphate (AcP), such as stationary-phase arrest and deletion of ackA, resulted in globally increased acetylation stoichiometry. Comparison to relative quantification under the same conditions validated our stoichiometry estimates at hundreds of sites, demonstrating the accuracy of our method. Similar to Sirtuin deacetylase 3 (SIRT3) in mitochondria, CobB suppressed acetylation to lower than median stoichiometry in WT, ptaΔ, and ackAΔ cells. Together, our results provide a detailed view of acetylation stoichiometry in E. coli and suggest an evolutionarily conserved function of Sirtuin deacetylases in suppressing low stoichiometry acetylation.
Article
Full-text available
Aspirin, or acetylsalicylic acid is widely used to control pain, inflammation and fever. Important to this function is its ability to irreversibly acetylate cyclooxygenases at active site serines. Aspirin has the potential to acetylate other amino-acid side-chains, leading to the possibility that aspirin-mediated lysine acetylation could explain some of its as-yet unexplained drug actions or side-effects. Using isotopically labeled aspirin-d3, in combination with acetylated lysine purification and LC-MS/MS, we identified over 12000 sites of lysine acetylation from cultured human cells. Although aspirin amplifies endogenous acetylation signals at the majority of detectable endogenous sites, cells tolerate aspirin mediated acetylation very well unless cellular deacetylases are inhibited. Although most endogenous acetylations are amplified by orders of magnitude, lysine acetylation site occupancies remain very low even after high doses of aspirin. This work shows that while aspirin has enormous potential to alter protein function, in the majority of cases aspirin-mediated acetylations do not accumulate to levels likely to elicit biological effects. These findings are consistent with an emerging model for cellular acetylation whereby stoichiometry correlates with biological relevance, and deacetylases act to minimize the biological consequences non-specific chemical acetylations.
Article
Full-text available
Post-translational modification of lysine residues by NƐ-acylation is an important regulator of protein function. Many large-scale protein acylation studies have assessed relative changes of lysine acylation sites after antibody enrichment using mass spectrometry-based proteomics. Although relative acylation fold-changes are important, this does not reveal site occupancy, or stoichiometry, of individual modification sites, which is critical to understand functional consequences. Recently, methods for determining lysine acetylation stoichiometry have been proposed based on ratiometric analysis of endogenous levels to those introduced after quantitative per-acetylation of proteins using stable isotope-labeled acetic anhydride. However, in our hands, we find that these methods can overestimate acetylation stoichiometries because of signal interferences when endogenous levels of acylation are very low, which is especially problematic when using MS1 scans for quantification. In this study, we sought to improve the accuracy of determining acylation stoichiometry using data-independent acquisition (DIA). Specifically, we use SWATH acquisition to comprehensively collect both precursor and fragment ion intensity data. The use of fragment ions for stoichiometry quantification not only reduces interferences but also allows for determination of site-level stoichiometry from peptides with multiple lysine residues. We also demonstrate the novel extension of this method to measurements of succinylation stoichiometry using deuterium-labeled succinic anhydride. Proof of principle SWATH acquisition studies were first performed using bovine serum albumin for both acetylation and succinylation occupancy measurements, followed by the analysis of more complex samples of E. coli cell lysates. Although overall site occupancy was low (<1%), some proteins contained lysines with relatively high acetylation occupancy. [Figure not available: see fulltext.]
Article
Full-text available
Nε-Lysine acetylation has been recognized as a ubiquitous regulatory posttranslational modification that influences a variety of important biological processes in eukaryotic cells. Recently, it has been realized that acetylation is also prevalent in bacteria. Bacteria contain hundreds of acetylated proteins, with functions affecting diverse cellular pathways. Still, little is known about the regulation or biological relevance of nearly all of these modifications. Here we characterize the cellular growth-associated regulation of the Bacillus subtilis acetylome. Using acetylation enrichment and quantitative mass spectrometry, we investigate the logarithmic and stationary growth phases, identifying over 2,300 unique acetylation sites on proteins that function in essential cellular pathways. We determine an acetylation motif, EK(ac)(D/Y/E), which resembles the eukaryotic mitochondrial acetylation signature, and a distinct stationary-phase-enriched motif. By comparing the changes in acetylation with protein abundances, we discover a subset of critical acetylation events that are temporally regulated during cell growth. We functionally characterize the stationary-phase-enriched acetylation on the essential shape-determining protein MreB. Using bioinformatics, mutational analysis, and fluorescence microscopy, we define a potential role for the temporal acetylation of MreB in restricting cell wall growth and cell diameter.
Article
The acetyltransferases CBP and p300 are multifunctional transcriptional co-activators. Here, we combined quantitative proteomics with CBP/p300-specific catalytic inhibitors, bromodomain inhibitor, and gene knockout to reveal a comprehensive map of regulated acetylation sites and their dynamic turnover rates. CBP/p300 acetylates thousands of sites, including signature histone sites and a multitude of sites on signaling effectors and enhancer-associated transcriptional regulators. Time-resolved acetylome analyses identified a subset of CBP/p300-regulated sites with very rapid (<30 min) acetylation turnover, revealing a dynamic balance between acetylation and deacetylation. Quantification of acetylation, mRNA, and protein abundance after CBP/p300 inhibition reveals a kinetically competent network of gene expression that strictly depends on CBP/p300-catalyzed rapid acetylation. Collectively, our in-depth acetylome analyses reveal systems attributes of CBP/p300 targets, and the resource dataset provides a framework for investigating CBP/p300 functions and for understanding the impact of small-molecule inhibitors targeting its catalytic and bromodomain activities. A comprehensive look at CBP/p300 acetylation reveals dynamic changes that shape the regulation of protein function and gene expression.
Article
Mapping the proteome Proteins function in the context of their environment, so an understanding of cellular processes requires a knowledge of protein localization. Thul et al. used immunofluorescence microscopy to map 12,003 human proteins at a single-cell level into 30 cellular compartments and substructures (see the Perspective by Horwitz and Johnson). They validated their results by mass spectroscopy and used them to model and refine protein-protein interaction networks. The cellular proteome is highly spatiotemporally regulated. Many proteins localize to multiple compartments, and many show cell-to-cell variation in their expression patterns. Presented as an interactive database called the Cell Atlas, this work provides an important resource for ongoing efforts to understand human biology. Science , this issue p. eaal3321 ; see also p. 806
Article
Functional characterization of lysine acetylation pathway requires the quantitative measurement of the modification abundance at the stoichiometry level. Here, we developed a systematic workflow for global untargeted identification of site-specific Lys acetylation stoichiometries in mammalian cells. Our strategy includes an optimized protocol for in vitro chemical labeling of unmodified lysine with stable isotope-encoded acetyl-NHS ester, deep proteomic profiling with high resolution mass spectrometer and a new software tool for quantitative analysis and stoichiometry determination. The workflow was validated using in vitro chemically labeled BSA and synthetic peptides with multiple Lys acetylation at various positions. In the proof-of-concept study, we applied the strategy to analyze the proteome of HeLa cells and determined the stoichiometries of over 600 acetylation sites with good reproducibility. Sodium butyrate treatment induced significant increase of acetylation stoichiometries in HeLa cells. Analysis of site-specific stoichiometry dynamics revealed the co-regulation of closely positioned acetylation sites on histone H3 and H4 upon treatment.