You are on page 1of 6

Journal of Thermal Biology 74 (2018) 71–76

Contents lists available at ScienceDirect

Journal of Thermal Biology


journal homepage: www.elsevier.com/locate/jtherbio

Roles for lysine acetyltransferases during mammalian hibernation T



Andrew N. Rouble, Liam J. Hawkins, Kenneth B. Storey
Institute of Biochemistry and Department of Biology, Carleton University, 1125 Colonel By Drive, Ottawa, ON K1S 5B6, Canada

A R T I C LE I N FO A B S T R A C T

Keywords: The thirteen-lined ground squirrel (Ictidomys tridecemlineatus) is a well-known model for studying hibernation.
Hibernation While in a torpid state, these animals globally suppress energy expensive processes, while supporting specialized
Acetyltransferases pathways necessary for survival. Lysine acetyltransferases (KATs) play a crucial role in modulating the ex-
Skeletal muscle pression and activity of a wide-variety of cellular pathways and processes, and therefore, may play a role during
Liver
hibernation when the cell is shifting to an energy conservative, cytoprotective state. Here we measured protein
Adipose tissue
levels of four KATs (CBP, PCAF, GCN5L2, HAT1), total histone acetyltransferase (HAT) activity, and the levels of
acetylation of histone H3 lysine 9 (H3K9ac), in multiple tissues across the torpor-arousal cycle. Our results show
a tissue-specific response of KATs, particularly in the adipose tissues where specific KATs (PCAF and GCN5L2),
HAT activity, and H3K9ac increased in the metabolically active BAT while HAT1, HAT activity and H3K9ac
decreased in WAT. Liver showed significant increases in the KAT PCAF whereas skeletal muscle had decreased
CBP and GCN5L2. Both liver and skeletal muscle showed no change in HAT activity and H3K9me3 increased in
muscle during torpor. Together, these results suggest KATs may play specialized roles in the different tissues of
the ground squirrel to contribute to the hibernator phenotype.

1. Introduction thermogenesis from uncoupled respiration in the specialized mi-


tochondria of brown adipose tissue (BAT). Remarkably, these animals
Some mammals have adapted to survive the winter months, when endure little to no damage to their tissues, or atrophy of their muscles
food resources become limited and temperatures drop below 0 °C, by (Andres-Mateos et al., 2013), despite prolonged periods of inactivity,
entering hibernation. During hibernation, these animals lower their rapid tissue reperfusion, and exposure to the production of high levels
metabolic rate for weeks to months at a time, with interspersed but of reactive oxygen species due to heightened oxygen consumption
brief periods of arousal where their metabolism temporarily returns to during arousal (Storey, 2010). Interestingly, it is generally accepted
euthermic levels (French, 1985). While there is considerable variety in that these extraordinary metabolic and protective adaptations are the
hibernation parameters (e.g. duration, feeding vs. non-feeding, range of result of the regulation of gene expression, rather than novel hiberna-
heterothermicity) depending on taxa and climatic region (Giroud et al., tion-specific genes (Carey et al., 2003). Therefore, the mechanisms of
2008; Nespolo et al., 2010; Weitten et al., 2016), during torpor, animals genetic regulation employed by hibernators have made these animals
such as the thirteen-lined ground squirrel (Ictidomys tridecemlineatus) particularly interesting subjects in the fields of gerontology, obesity,
display drastic reductions in body temperature (Tb), heart rate, diabetes, and neurodegenerative diseases (Carey et al., 2003; Härtig
breathing rate, and organ perfusion rates (Bullard and Funkhouser, et al., 2007; Martin, 2008; Storey, 2010; Wu and Storey, 2016).
1962; Carey et al., 2003; Frerichs et al., 1994). Congruent with these The molecular regulation of hibernation is in large part dependent
physiological changes are alterations to energy metabolism pathways on the reversible post-translational modification of proteins (Bell and
that result in a shift away from carbohydrate metabolism towards lipid Storey, 2017; Rouble and Storey, 2015; Storey, 2010; Tessier et al.,
catabolism, particularly in the liver, and fatty acid oxidation of lipid 2017). We have previously investigated the role of reversible protein
reserves in white adipose tissue (WAT) accumulated during the summer acetylation in this context, with specific focus on the expression and
months (Carey et al., 2003; Dark, 2005). These lipid deposits are es- activity of NAD+-dependent SIRT deacetylases (Rouble and Storey,
sential in placental hibernators during periods of arousal, when in- 2015) and histone deacetylases (HDACs) (Morin and Storey, 2006). The
creases in oxygen consumption and fatty acid oxidation fuel counter-parts to these deacetylase enzymes are lysine acetyltransferases

Abbreviations: KAT, lysine acetyltransferase; HAT, histone acetyltransferase; H3K9ac, histone H3 lysine 9 acetylation; BAT, brown adipose tissue; WAT, white adipose tissue; Tb, body
temperature; HDAC, histone deacetylase; IA, interbout arousal; EC, euthermic in cold room; EN, entrance into torpor; ET, early torpor; LT, late torpor; EA, early arousal

Corresponding author.
E-mail address: kenneth_storey@carleton.ca (K.B. Storey).

https://doi.org/10.1016/j.jtherbio.2018.03.013
Received 9 February 2018; Received in revised form 13 March 2018; Accepted 13 March 2018
Available online 15 March 2018
0306-4565/ © 2018 Elsevier Ltd. All rights reserved.
A.N. Rouble et al. Journal of Thermal Biology 74 (2018) 71–76

(KATs), which function to add acetyl-moieties to lysine residues. The


addition of acetyl-groups to target proteins can alter important mole-
cular parameters, such as DNA-binding affinities, enzymatic activities,
protein stability, target specificity, and complex formation (Carrozza
et al., 2003; Chen et al., 2001; Spange et al., 2009). KATs also enable
the histone-acetylation-dependent epigenetic remodeling of chromatin,
thereby controlling the activation of transcription at specific loci
(Carrozza et al., 2003; Lee and Workman, 2007; Verdone et al., 2006).
Four of the best-characterized KATs are CREB Binding Protein
(CBP), P300/CBP-associated factor (PCAF), General Control of Amino
Acid Synthesis Protein 5-Like 2 (GCN5L2), and Histone
Acetyltransferase 1 (HAT1). The functions of these proteins is integral
in processes such as cell death, DNA repair, proliferation, metabolism,
and general transcriptional activation (Chen et al., 2001; Spange et al.,
2009; Xiong and Guan, 2012). Because of these important roles, these
KATs are of great interest to researchers seeking novel approaches to
the treatment of various diseases, and similarly, as potential targets for
the regulation processes that allow hibernating mammals to undergo
metabolic rate depression during torpor. In this study, we explore KATs
in the context of a mammalian hibernator, the thirteen-lined ground
squirrel. Protein levels of CBP, PCAF, GCN5L2, and HAT1 were mea-
Fig. 1. Relative protein expression of CBP, PCAF, GCN5L2 and HAT1 in liver of I. tride-
sured over the five standardized time points of the torpor-arousal cycle cemlineatus over the torpor-arousal cycle. Representative protein bands are shown for
(Carey et al., 2003; Storey, 2010) with a euthermic control in four selected sampling points (labelled to the left and right of the blots). Histogram shows
tissues: liver, skeletal muscle, BAT, and WAT. These tissues were se- mean standardized band densities ( ± SEM, n = 4). Protein bands were standardized
lected due to their previously mentioned physiology, metabolism, and against the summed intensity of a group of Coomassie-stained protein bands from the
resilience during hibernation. Additionally, total histone acetyl- same sample lane. Data were analyzed using a one-way ANOVA with a post hoc Dunnett's
test. The symbol * indicates significant difference from the respective EC control,
transferase (HAT) activities were measured, comparing euthermic ani-
p < 0.05.
mals to those in deep-torpor. Finally, the levels of acetylated histone H3
lysine 9 (H3K9ac), a KAT-targeted histone residue, were measured
across the torpor-arousal cycle. The results reinforce previous studies
on the role for reversible protein acetylation in the context of mam-
malian hibernation (Rouble and Storey, 2015).

2. Results

2.1. Analysis of KAT protein levels over the torpor-arousal cycle

Relative protein levels of select KATs (CBP, PCAF, GCN5L2, and


HAT1) were measured in the liver (Fig. 1), skeletal muscle (Fig. 2), BAT
(Fig. 3), and WAT (Fig. 4) of the ground squirrel over the six time-points
of the torpor arousal cycle. In liver, protein levels of CBP were sig-
nificantly reduced to 0.4 ± 0.08 during interbout arousal (IA) as
compared to values for animals that were euthermic in the cold room
(EC), while levels of PCAF were significantly elevated over controls
during four stages: entrance into torpor (EN), early torpor (ET), late
torpor (LT) and early arousal (EA) (by 4.7 ± 0.2, 2.2 ± 0.09,
3.0 ± 0.2, and 4.6 ± 0.3-fold, respectively). Protein levels of GCN5L2
were also significantly decreased to 0.6 ± 0.03 during LT when com-
pared to EC, and subsequently increased during EA by 1.4 ± 0.2-fold
over controls. No significant fluctuations were observed for protein Fig. 2. Relative protein expression of CBP, PCAF, GCN5L2 and HAT1 in skeletal muscle of
I. tridecemlineatus over the torpor-arousal cycle. All other information as in Fig. 1.
levels of HAT1 in liver over the torpor-arousal cycle.
In skeletal muscle, CBP protein levels were significantly reduced
during LT to 0.3 ± 0.06 of the corresponding EC values, while levels of did not change significantly at any of the sampled time points.
GCN5L2 were also significantly lowered during EN, LT, EA and IA (to
0.7 ± 0.03, 0.6 ± 0.03, 0.7 ± 0.1, and 0.7 ± 0.06 of EC, respec- 2.2. Analysis of total HAT activity
tively), but not during ET. Protein levels of PCAF and HAT1 did not
change significantly over torpor-arousal. Total relative HAT activity was measured in the total soluble protein
In BAT, protein levels of PCAF were significantly elevated during EN fraction of liver, skeletal muscle, BAT, and WAT, comparing EC and LT
by 1.9 ± 0.1-fold over EC, while levels of GCN5L2 were also elevated stages (Fig. 5). In liver and skeletal muscle, relative HAT activity did
during LT by 1.4 ± 0.04-fold before significantly decreasing during IA not change significantly during LT as compared to EC control (Liver EC
to 0.7 ± 0.01 of EC. Protein levels of HAT1 were also significantly – 2869 ± 743 ng/h/mg, LT – 2377 ± 297 ng/h/mg; Muscle EC –
reduced during ET and IA to 0.6 ± 0.02 and 0.4 ± 0.05 of EC values, 749 ± 39 ng/h/mg, LT – 808 ± 105 ng/h/mg). In contrast, relative
respectively. In contrast, levels of CBP did not fluctuate significantly HAT activity in BAT was significantly elevated during LT by 2.0 ± 0.2-
over the course of the torpor-arousal cycle. fold as compared to EC (EC – 2407 ± 581 ng/h/mg, LT –
In WAT, protein levels of HAT1 decreased significantly during LT to 4754 ± 497 ng/h/mg), whereas total KAT activity in WAT was sig-
0.5 ± 0.04 of the EC value, whereas levels of CBP, PCAF and GCN5L2 nificantly reduced during LT to 0.5 ± 0.07 of the EC value (EC –

72
A.N. Rouble et al. Journal of Thermal Biology 74 (2018) 71–76

Fig. 5. Total relative HAT activity in liver, skeletal muscle, brown adipose tissue and
white adipose tissue of I. tridecemlineatus comparing euthermic control (EC) and late
torpor (LT) points of the torpor-arousal cycle. Histograms show means ± SEM, n = 4.
Data were analyzed using the Student's t-test. The symbol * indicates significant differ-
ence from the respective EC control, p < 0.05.

Fig. 3. Relative protein expression of CBP, PCAF, GCN5L2 and HAT1 in brown adipose
tissue of I. tridecemlineatus over the torpor-arousal cycle. All other information as in Fig. 1.

Fig. 6. Relative protein expression of histone H3 acetylated at lysine 9 (H3K9ac) in liver,


skeletal muscle, brown adipose tissue and white adipose tissue of I. tridecemlineatus over
the torpor-arousal cycle. All other information as in Fig. 1.

Fig. 4. Relative protein expression of CBP, PCAF, GCN5L2 and HAT1 in white adipose
tissue of I. tridecemlineatus over the torpor-arousal cycle. All other information as in Fig. 1. action in transcriptional complexes (Chan and La Thangue, 2001; Nagy
and Tora, 2007; Roth et al., 2001). Since transcription is tightly con-
96 ± 14 ng/h/mg, LT – 46 ± 7 ng/h/mg). trolled in hibernating mammals, it may be the case that KATs such as
CBP, PCAF, GNC5L2, and HAT1 are also regulated or involved in the
regulation of this energy intensive process. Our previous studies have
2.3. Analysis of acetylation status of histone H3 lysine 9 during hibernation
identified possible roles for HDACs and SIRT-deacetylases in metabolic
rate depression and cellular protective pathways during hibernation
Relative protein levels of histone H3 acetylated at lysine 9 (H3K9ac)
(Morin and Storey, 2006; Rouble and Storey, 2015). However, given
(Fig. 6) were measured in ground squirrel liver, muscle, BAT and WAT,
that KATs have never previously been characterized in hibernators,
over the six sampling-points of the torpor-arousal cycle. Relative
their regulation in the context of this form of metabolic suppression has
amounts of H3K9ac protein increased significantly in BAT during ET
been unknown, representing a major gap in knowledge within the field
and LT (by 2.4 ± 0.2 and 2.2 ± 0.2-fold, respectively) and in muscle
of hibernation research. With the goal of filling this gap, this study has
during ET (by 1.8 ± 0.1-fold) as compared to the respective EC con-
attempted to provide an initial characterization of KAT involvement in
trols. In WAT, H3K9 levels decreased significantly during ET to
hibernation, and succeeds in providing evidence to suggest a role for
0.4 ± 0.02 before increasing by 2.2 ± 0.2-fold during IA, as com-
these enzymes within this context.
pared to EC. In liver, no significant fluctuations in H3K9 levels were
observed.
3.1. Hibernation and KAT protein levels
3. Discussion
Our results show that, of the four KATs considered, these enzymes
KATs have been implicated in the regulation of a wide variety of are differentially expressed in a tissue-specific manner at various points
cellular processes through their modification of histones and direct during hibernation, but do not exhibit an overall pattern of universal

73
A.N. Rouble et al. Journal of Thermal Biology 74 (2018) 71–76

protein suppression or activation (Figs. 1–4). This observation would demonstrated significant fluctuations in HAT activity. In BAT, HAT
suggest that, in these animals, each of the studied factors likely serves activity levels doubled during torpor as compared to control, thereby
some distinct role(s) at different times, and in different tissues, further supporting a role for increased HAT function in this tissue
throughout hibernation. For example, PCAF protein expression was during hypometabolism. As mentioned, increased HAT activity in BAT
strongly enhanced in liver from EN through to EA (returning to eu- might reflect the need for this tissue to maintain the expression of
thermic levels during the interbout period), while this same pattern was certain pathways involved in the thermoregulatory response, perhaps
not observed in any other tissue. This suggests that the increased ex- via acetylation-induced increases in the transcriptional activation of
pression of PCAF serves a specific role in liver during torpor that does specific genes. In contrast, HAT activity in WAT was reduced by half
not appear to be required in other tissues. Interestingly, under condi- during torpor, suggesting that a reduction in the function of these en-
tions that inhibit glycolysis, PCAF is known to acetylate and reduce the zymes is required in this tissue during metabolic depression. Interest-
activity of pyruvate kinase (Lv et al., 2011), a major glycolytic enzyme ingly, the activity fluctuations in both tissues seem to correlate with
that is also suppressed by RPP during hibernation (Storey and Storey, some of the observed protein data – in BAT, GCN5L2 levels increase
2010). The increased expression of PCAF in liver throughout hiberna- with total HAT activity during torpor, while levels of HAT1 decrease
tion may therefore contribute to the inhibition of certain metabolic with total activity in WAT, possibly supporting the notion that these
processes such as glycolysis, in a manner analogous to the increased increases/decreases in protein levels make an actual contribution to the
activities of specific kinases/phosphatases that also occur during this total measurable acetyltransferase activity within the tissues. However,
time in the liver. Protein levels of CBP also demonstrated a tissue-spe- given that these measurements only accounted for total HAT activity –
cific response, most notably being suppressed during late torpor only in and not the activity of specific KATs – no conclusions can be made
muscle tissue. Given CBP's role as a diverse transcriptional co-activator regarding the relationship between relative changes in the protein ex-
(Chan and La Thangue, 2001; Kalkhoven, 2004), this fluctuation may pression of one or two enzymes, and total HAT activity. This concept
serve an important purpose in the regulation of the metabolic sup- also applies to the lack of change in activity observed in muscle and
pression in hibernator muscle. This is because its presence and acetyl- liver – although protein fluctuations occurred in both tissues, ob-
transferase activity in certain transcriptional complexes is responsible servable changes in total activity will not necessarily follow, even if the
and necessary for the regulation of many different genes; therefore, a activities of specific KATs do change. Regardless of the specificity of the
decrease in CBP expression could contribute to the widespread tran- measurements, however, the current data support a role for overall
scriptional suppression that is characteristic of the torpid state. In fact, enhanced HAT activity in BAT and decreased HAT activity in WAT
this proposed function would also complement current evidence that during torpor, thereby providing further evidence to implicate these
suggests an epigenetic contribution to global transcriptional arrest. In enzymes in the regulation of hibernation.
hibernator muscle, the acetylation of histone H3 at lysine residue 23
(H3K23) is known to be significantly reduced during torpor, which is 3.3. Hibernation responsive acetyl-histone levels
consistent with the notion of suppressed transcriptional activity at loci
associated with this histone modification (Morin and Storey, 2006). Transcriptional control through KAT-mediated acetylation of his-
This change is also correlated with an increase in the expression levels tone residues has been shown to be integral to the activation of
of several HDACs which target this residue, in addition to levels of total countless genes and the regulation of crucial cellular processes
HDAC activity which was demonstrated as a proof of concept in (Carrozza et al., 2003; Jenuwein and Allis, 2001; Verdone et al., 2006).
Hawkins and Storey (2017) and Morin and Storey (2006). CBP also Thus, to more fully characterize KATs in the context of hibernation, we
targets this residue as a substrate for its acetyltransferase activity measured the acetylation levels of KAT-targeted histone H3 lysine 9
(Henry et al., 2013). Thus, the observed decrease in CBP expression (H3K9ac) across the torpor-arousal cycle (Fig. 6). Acetylation of H3K9
during late torpor may compliment the increase in HDAC activity to is generally associated with active transcription, and multiple KATs
promote the deacetylation of certain histone residues, and thereby measured in this study (PCAF, GCN5L2 and CBP) target this residue
promote transcriptional suppression. This may also be true for GCN5L2, (Karmodiya et al., 2012). Similar to the measured KAT protein levels
the protein levels of which were also significantly reduced during late and HAT activities, protein levels of H3K9ac showed tissue-specific
torpor in muscle. GCN5L2 targets H3K23 for acetylation (Grant et al., changes, suggesting that transcriptional activation/deactivation of
1999), so downregulation of GCN5L2 may contribute to the reduced certain processes by differential histone acetylation may occur at var-
acetylation of histone H3. Like CBP, GCN5L2 is also a major tran- ious points throughout the torpor-arousal cycle. Notably, H3K9ac was
scriptional co-activator in a non-epigenetic sense (Nagy and Tora, significantly elevated in BAT during torpor, which is consistent with the
2007), so its decreased expression in muscle and liver may also gen- transcriptional activation of gene programs regulated by this residue
erally reflect the global suppression of transcription during torpor. In being a part of this tissue's response to torpor. In fact, Evidence exists to
contrast, the enhanced expression of GCN5L2 at the same time in BAT suggest that the regulation of uncoupling protein-1 (UCP1, the main
could suggest an increase in the transcriptional activation of GCN5L2 protein responsible for uncoupled respiration and non-shivering ther-
target-genes, which could be functional in this tissue given its potential mogenesis in the mitochondria of BAT) is controlled by histone H3
role in regulating thermogenesis when Tb drops below acceptable limits acetylation, whereby decreased levels of the modification are asso-
during torpor (Boyer and Barnes, 1999), or to initiate arousal (Nizielski ciated with reductions in ucp1 gene expression (Kiskinis et al., 2007).
et al., 1989). While the actual functions of these and other observed Given that UCP1 expression is absolutely integral to the thermo-
changes in KAT protein expression cannot be conclusively determined regulatory function of BAT (Golozoubova et al., 2001), and is therefore
by the current study, the fact that such changes do occur is evidence to indispensable for the survival of the hibernating mammal, the tran-
suggest that these enzymes likely serve roles in the regulation of the scriptional activation of the ucp1 gene by H3K9 acetylation in BAT
various processes implicated during hibernation. during torpor would be unsurprising. Similarly, the significant fluc-
tuation in H3K9 acetylation between early torpor and interbout arousal
3.2. Histone acetyltransferase activity through the torpor-arousal cycle in WAT may reflect general changes in transcriptional activity that
occur over this period (i.e. suppression during the initial metabolic
To further characterize the possible function of KATs in the context decline, followed by strong reactivation during arousal). Interestingly,
of the hibernator, total HAT activity (HAT enzymatic activity is carried the results in BAT correlate with the observed increases in HAT activity
out by KATs) in the four tissues was compared between euthermic (EC) and GCN5L2 protein levels that also occur in BAT at this time, possibly
and late torpor (LT) stages (Fig. 5). While no change in HAT activity reflecting the expected change in downstream substrate acetylation that
occurred in liver or muscle between these stages, the adipose tissues should occur with fluctuations in HAT activity/expression. While the

74
A.N. Rouble et al. Journal of Thermal Biology 74 (2018) 71–76

exact functions of the observed changes in H3K9 acetylation remain Coomassie blue (0.25% w/v Coomassie brilliant blue, 7.5% v/v acetic
unknown, the current data support the idea that epigenetic mechanisms acid, 50% methanol) for loading standardization.
likely contribute to the adaptation involved in the torpor-arousal cycle.
4.3. Lysine acetyltransferase activity assay
3.4. Conclusion
Total histone acetyltransferase (HAT) activity during EC and LT was
This study serves as the first known investigation to provide sig- assayed in liver, skeletal muscle, BAT, and WAT total protein extracts
nificant evidence to suggest that the differential expression of KATs using the EpiQuick HAT Activity/Inhibition Assay Kit (Epigentek, P-
may be characteristic of the hibernation phenotype. Indeed, the results 4003) as per the manufacturer's instructions. Samples were prepared as
discussed herein identify fluctuations in the protein levels of four of the above with the exception that extracts were not mixed with 2 × SDS
best-studied KATs in the literature, changes in HAT activity, and dif- loading buffer. Briefly, in each assay well, 50 µL of 1:50 HAT substrate
ferential acetylation of a downstream KAT histone target, at various (supplied by the manufacturer) was incubated at room temperature for
points throughout the torpor-arousal cycle and in a tissue-specific 45 min, which was then aspirated and each well was washed with
manner. Some of these changes also appear to correlate. For example, 150 µL of wash buffer (supplied by the manufacturer) three times.
during torpor in BAT, the increased expression of the major transcrip- Then, added to each well was 2 µL of protein extracts, 26 µL of HAT
tional co-activator GCN5L2 occurs concurrently with increases in HAT assay buffer (supplied by the manufacturer), and 2 µL of acetyl CoA
activity and H3K9 acetylation, all of which are changes that point to- (1:20 v/v from 30 mM stock in HAT assay buffer), which were then
wards enhanced transcriptional activation via KAT-mediated regula- incubated for 60 min at 37 °C. The wells were then washed three times
tion. While specific impacts on the hibernation phenotype by these as above and 50 µL of capture antibody (supplied by the manufacturer)
proteins will need to be explored further, these data likely represent was added and incubated for 60 min at room temperature on an orbital
further examples of the widespread function of reversible protein shaker. The wells were wash four times, and 50 µL of detection anti-
acetylation in the regulation of diverse cellular processes, and suggests body (1:1000, supplied by the manufacturer) was added and incubated
its importance to hibernator biology. for 30 min at room temperature on an orbital shaker. Each well was
then washed five times and 100 µL of developer solution was added to
4. Materials and methods each well and incubated for 10 min at room temperature on an orbital
shaker in the dark. 50 µL of stop solution was then added to each well
4.1. Animal care and treatment and the absorbance of each well was read at 450 nm using a Powerwave
HT spectrophotometer (BioTek). Three wells with additional assay
Animal experiments were performed as previously described buffer instead of protein extracts were run during the assay to act as
(McMullen and Hallenbeck, 2010; Rouble et al., 2013) by Dr. J.M. negative controls as per the manufacturer's instructions.
Hallenbeck and were approved by the Animal Care and Use Committee
of the National Institute of Neurological Disorders and Stroke (NIH; 4.4. Quantification and statistics
animal protocol no. ASP 1223– 05). Ictidomys tridecemlineatus were used
in this study and are small mammalian winter hibernators that enter Band densities on chemiluminescent immunoblots were quantified
bouts of torpor lasting days to weeks with periods of rewarming and using GeneTools (Syngene, Frederick, MD). Band densities were stan-
arousal that can last ~ 24 h. Animals were sacrificed throughout the dardized against the summed intensity of Coomassie stained protein
torpor-arousal cycle, and liver, skeletal muscle, BAT, and WAT were bands in the same lane and then normalized to their respective EC
quickly excised and frozen in liquid nitrogen. Animals were sampled at condition. Data are expressed as mean ± SEM, n = 4. Statistical ana-
the following time points: (1) euthermic in the cold room (EC) main- lysis of the data was performed by a one-way ANOVA with a Dunnett's
tained at 4 °C, as previously described (McMullen and Hallenbeck, post-hoc test (p < 0.05) to correct for multiple comparisons using
2010) – these animals had not entered torpor for at least 72 h and had SigmaPlot 12 statistical package software (Systat Software Inc., San
Tb of 36–37 °C. (2) Entrance into torpor (EN) – these animals have Jose, CA, USA). HAT activity assays were corrected using negative
shown a decline in Tb (18–31 °C) and have begun to enter torpor. (3) control wells, data are expressed as mean ± SEM, n = 4 and normal-
Early torpor (ET) – animals that are in torpor for 24 h with a constant Tb ized to the EC samples. Statistical analysis of HAT activity assay results
of 5–8 °C. (4) Late torpor (LT) – these animals had been in torpor for at was performed by Student's t-tests (p < 0.05).
least 5 days with a constant Tb of 5–8 °C. (5) Early arousal (EA) – these
animals show a rising Tb and are sampled when Tb was 9–12 °C. (6) Acknowledgments
Interbout arousal (IA) – these animals have arisen from torpor and their
Tb has return to euthermic levels (~ 37 °C) for ~ 18 h. Tissues were The authors thank Dr. J.M. Hallenbeck and Dr. D.C. McMullen
transported on dry ice to Carleton University and stored at − 80 °C until (NINDS, NIH, Bethesda) for providing the tissue samples for this study
use. and Jan Storey for assistance in the editing of the manuscript.

4.2. Protein extraction and immunoblotting Funding

Total protein extraction and immunoblotting was performed as This work was supported by a Discovery grant (Grant # 6793) from
previously described (Rouble et al., 2013). Membranes were blocked the Natural Sciences and Engineering Research Council (NSERC) of
with milk (2.5–5%, 20–30 min) or polyvinyl alcohol (1 mg/mL, Canada. ANR held a NSERC CGSM Scholarship, and KBS holds the
30–70 kDa PVA, 45–60 s) in tris-buffered saline with Tween-20 (TBST). Canada Research Chair in Molecular Physiology.
Targets were probed with antibodies for HAT1 (Genetex, GTX110643),
GCN5L2 (Cell Signaling, #3305), PCAF (Cell Signaling, #3378), CBP Conflicts of interest
(Cell Signaling, #7389), or H3K9ac (Cell Signaling, #9649)
(1:1000–12000 v/v in TBST) overnight at 4 °C. Membranes were then The authors report no conflicts of interest.
probed with HRP-conjugated anti-rabbit secondary antibodies (1:1500-
8000 v/v in TBST, 30–60 min) at room temperature and visualized with References
chemiluminescence using the ChemiGenius Bio Imaging System (Syn-
gene, Frederick, MD, USA). Membranes were then stained with Andres-Mateos, E., Brinkmeier, H., Burks, T.N., Mejias, R., Files, D.C., Steinberger, M.,

75
A.N. Rouble et al. Journal of Thermal Biology 74 (2018) 71–76

Soleimani, A., Marx, R., Simmers, J.L., Lin, B., Finanger Hedderick, E., Marr, T.G., a subset of inactive inducible promoters in mouse embryonic stem cells. BMC Genom.
Lin, B.M., Hourdé, C., Leinwand, L.A., Kuhl, D., Föller, M., Vogelsang, S., Hernandez- 13, 424. http://dx.doi.org/10.1186/1471-2164-13-424.
Diaz, I., Vaughan, D.K., Alvarez de la Rosa, D., Lang, F., Cohn, R.D., 2013. Activation Kiskinis, E., Hallberg, M., Christian, M., Olofsson, M., Dilworth, S.M., White, R., Parker,
of serum/glucocorticoid-induced kinase 1 (SGK1) is important to maintain skeletal M.G., 2007. RIP140 directs histone and DNA methylation to silence Ucp1 expression
muscle homeostasis and prevent atrophy. EMBO Mol. Med. 5, 80–91. http://dx.doi. in white adipocytes. EMBO J. 26, 4831–4840. http://dx.doi.org/10.1038/sj.emboj.
org/10.1002/emmm.201201443. 7601908.
Bell, R.A.V., Storey, K.B., 2017. Purification and characterization of skeletal muscle Lee, K.K., Workman, J.L., 2007. Histone acetyltransferase complexes: one size doesn’t fit
pyruvate kinase from the hibernating ground squirrel, Urocitellus richardsonii: po- all. Nat. Rev. Mol. Cell Biol. 8, 284–295. http://dx.doi.org/10.1038/nrm2145.
tential regulation by posttranslational modification during torpor. Mol. Cell. Lv, L., Li, D., Zhao, D., Lin, R., Chu, Y., Zhang, H., Zha, Z., Liu, Y., Li, Z., Xu, Y., Wang, G.,
Biochem. 1–12. http://dx.doi.org/10.1007/s11010-017-3192-9. Huang, Y., Xiong, Y., Guan, K.-L., Lei, Q.-Y., 2011. Acetylation targets the M2 isoform
Boyer, B.B., Barnes, B.M., 1999. Molecular and metabolic aspects of mammalian hi- of pyruvate kinase for degradation through chaperone-mediated autophagy and
bernation: expression of the hibernation phenotype results from the coordinated promotes tumor growth. Mol. Cell 42, 719–730. http://dx.doi.org/10.1016/j.molcel.
regulation of multiple physiological and molecular events during preparation for and 2011.04.025.
entry into torpor. Bioscience 49, 713–724. http://dx.doi.org/10.2307/1313595. Martin, S.L., 2008. Mammalian hibernation: a naturally reversible model for insulin re-
Bullard, R.W., Funkhouser, G.E., 1962. Estimated regional blood flow by rubidium 86 sistance in man? Diabetes Vasc. Dis. Res. 5, 76–81. http://dx.doi.org/10.3132/dvdr.
distribution during arousal from hibernation. Am. J. Physiol. 203, 266–270. http:// 2008.013.
dx.doi.org/10.1152/ajplegacy.1962.203.2.266. McMullen, D.C., Hallenbeck, J.M., 2010. Regulation of Akt during torpor in the hi-
Carey, H.V., Andrews, M.T., Martin, S.L., 2003. Mammalian hibernation: cellular and bernating ground squirrel, Ictidomys tridecemlineatus. J. Comp. Physiol. B. 180,
molecular responses to depressed metabolism and low temperature. Physiol. Rev. 83, 927–934. http://dx.doi.org/10.1007/s00360-010-0468-8.
1153–1181. http://dx.doi.org/10.1152/physrev.00008.2003. Morin, P., Storey, K.B., 2006. Evidence for a reduced transcriptional state during hi-
Carrozza, M.J., Utley, R.T., Workman, J.L., Côté, J., 2003. The diverse functions of his- bernation in ground squirrels. Cryobiology 53, 310–318. http://dx.doi.org/10.1016/
tone acetyltransferase complexes. Trends Genet. 19, 321–329. http://dx.doi.org/10. j.cryobiol.2006.08.002.
1016/S0168-9525(03)00115-X. Nagy, Z., Tora, L., 2007. Distinct GCN5/PCAF-containing complexes function as co-ac-
Chan, H.M., La Thangue, N.B., 2001. p300/CBP proteins: HATs for transcriptional bridges tivators and are involved in transcription factor and global histone acetylation.
and scaffolds. J. Cell Sci. 114, 2363–2373. http://dx.doi.org/10.1007/BF00692921. Oncogene 26, 5341–5357. http://dx.doi.org/10.1038/sj.onc.1210604.
Chen, H., Tini, M., Evans, R.M., 2001. HATs on and beyond chromatin. Curr. Opin. Cell Nespolo, R.F., Verdugo, C., Cortés, P.A., Bacigalupe, L.D., 2010. Bioenergetics of torpor in
Biol. 13, 218–224. http://dx.doi.org/10.1016/S0955-0674(00)00200-3. the microbiotherid marsupial, monito del monte (Dromiciops gliroides): the role of
Dark, J., 2005. Annual lipid cycles in hibernators: integration of physiology and behavior. temperature and food availability. J. Comp. Physiol. B 180, 767–773. http://dx.doi.
Annu. Rev. Nutr. 25, 469–497. http://dx.doi.org/10.1146/annurev.nutr.25.050304. org/10.1007/s00360-010-0449-y.
092514. Nizielski, S.E., Billington, C.J., Levine, A.S., 1989. Brown fat GDP binding and circulating
French, A.R., 1985. Allometries of the durations of torpid and euthermic intervals during metabolites during hibernation and arousal. Am. J. Physiol. 257, R536–R541. http://
mammalian hibernation: a test of the theory of metabolic control of the timing of dx.doi.org/10.1152/ajpregu.1989.257.3.R536.
changes in body temperature. J. Comp. Physiol. B. 156, 13–19. http://dx.doi.org/10. Roth, S.Y., Denu, J.M., Allis, C.D., 2001. Histone acetyltransferases. Annu. Rev. Biochem.
1007/BF00692921. 70, 81–120. http://dx.doi.org/10.1146/annurev.biochem.70.1.81.
Frerichs, K.U., Kennedy, C., Sokoloff, L., Hallenbeck, J.M., 1994. Local cerebral blood Rouble, A.N., Hefler, J., Mamady, H., Storey, K.B., Tessier, S.N., 2013. Anti-apoptotic
flow during hibernation, a model of natural tolerance to “cerebral ischemia”. J. signaling as a cytoprotective mechanism in mammalian hibernation. PeerJ 1, e29.
Cereb. Blood Flow. Metab. 14, 193–205. http://dx.doi.org/10.1038/jcbfm.1994.26. http://dx.doi.org/10.7717/peerj.29.
Giroud, S., Blanc, S., Aujard, F., Bertrand, F., Gilbert, C., Perret, M., 2008. Chronic food Rouble, A.N., Storey, K.B., 2015. Characterization of the SIRT family of NAD+-dependent
shortage and seasonal modulations of daily torpor and locomotor activity in the grey protein deacetylases in the context of a mammalian model of hibernation, the thir-
mouse lemur (Microcebus murinus). Am. J. Physiol. Regul. Integr. Comp. Physiol. 294, teen-lined ground squirrel. Cryobiology 71, 334–343. http://dx.doi.org/10.1016/j.
R1958–R1967. http://dx.doi.org/10.1152/ajpregu.00794.2007. cryobiol.2015.08.009.
Golozoubova, V., Hohtola, E., Matthias, A., Jacobsson, A., Cannon, B., Nedergaard, J., Spange, S., Wagner, T., Heinzel, T., Krämer, O.H., 2009. Acetylation of non-histone
2001. Only UCP1 can mediate adaptive nonshivering thermogenesis in the cold. proteins modulates cellular signalling at multiple levels. Int. J. Biochem. Cell Biol. 41,
FASEB J. 15, 2048–2050. http://dx.doi.org/10.1096/fj.00-0536fje. 185–198. http://dx.doi.org/10.1016/j.biocel.2008.08.027.
Grant, P.A., Eberharter, A., John, S., Cook, R.G., Turner, B.M., Workman, J.L., 1999. Storey, K.B., 2010. Out cold: biochemical regulation of mammalian hibernation - a mini-
Expanded lysine acetylation specificity of Gcn5 in native complexes. J. Biol. Chem. review. Gerontology 56, 220–230. http://dx.doi.org/10.1159/000228829.
274, 5895–5900. http://dx.doi.org/10.1074/jbc.274.9.5895. Storey, K.B., Storey, J.M., 2010. Metabolic rate depression: the biochemistry of mam-
Hawkins, L.J., Storey, K.B., 2017. Improved high-throughput quantification of lumines- malian hibernation. Adv. Clin. Chem. 52, 77–108. http://dx.doi.org/10.1016/S0065-
cent microplate assays using a common Western-blot imaging system. MethodsX 4, 2423(10)52003-1.
413–422. http://dx.doi.org/10.1016/j.mex.2017.10.006. Tessier, S.N., Luu, B.E., Smith, J.C., Storey, K.B., 2017. The role of global histone post-
Härtig, W., Stieler, J., Boerema, A.S., Wolf, J., Schmidt, U., Weissfuss, J., Bullmann, T., translational modifications during mammalian hibernation. Cryobiology 75, 28–36.
Strijkstra, A.M., Arendt, T., 2007. Hibernation model of tau phosphorylation in http://dx.doi.org/10.1016/j.cryobiol.2017.02.008.
hamsters: selective vulnerability of cholinergic basal forebrain neurons - implications Verdone, L., Agricola, E., Caserta, M., Di Mauro, E., 2006. Histone acetylation in gene
for Alzheimer's disease. Eur. J. Neurosci. 25, 69–80. http://dx.doi.org/10.1111/j. regulation. Brief. Funct. Genom. Prote. 5, 209–221. http://dx.doi.org/10.1093/bfgp/
1460-9568.2006.05250.x. ell028.
Henry, R.A., Kuo, Y.-M., Andrews, A.J., 2013. Differences in specificity and selectivity Weitten, M., Oudart, H., Habold, C., 2016. Maintenance of a fully functional digestive
between CBP and p300 acetylation of histone H3 and H3/H4. Biochemistry 52, system during hibernation in the European hamster, a food-storing hibernator. Comp.
5746–5759. http://dx.doi.org/10.1021/bi400684q. Biochem. Physiol. A. Mol. Integr. Physiol. 193, 45–51. http://dx.doi.org/10.1016/j.
Jenuwein, T., Allis, C.D., 2001. Translating the histone code. Science 293, 1074–1080. cbpa.2016.01.006.
http://dx.doi.org/10.1126/science.1063127. Wu, C.-W., Storey, K.B., 2016. Life in the cold: links between mammalian hibernation and
Kalkhoven, E., 2004. CBP and p300: HATs for different occasions. Biochem. Pharmacol. longevity. Biomol. Concepts 7, 41–52. http://dx.doi.org/10.1515/bmc-2015-0032.
68, 1145–1155. http://dx.doi.org/10.1016/j.bcp.2004.03.045. Xiong, Y., Guan, K.-L., 2012. Mechanistic insights into the regulation of metabolic en-
Karmodiya, K., Krebs, A.R., Oulad-Abdelghani, M., Kimura, H., Tora, L., 2012. H3K9 and zymes by acetylation. J. Cell Biol. 198, 155–164. http://dx.doi.org/10.1083/jcb.
H3K14 acetylation co-occur at many gene regulatory elements, while H3K14ac marks 201202056.

76

You might also like