Metabolic Regulation of the Epitranscriptome - ACS Chemical Biology

Jan 17, 2019 - An emergent theme in cancer biology is that dysregulated energy metabolism may directly influence oncogenic gene expression. This is du...
2 downloads 0 Views 1MB Size
Subscriber access provided by EDINBURGH UNIVERSITY LIBRARY | @ http://www.lib.ed.ac.uk

Review

Metabolic Regulation of the Epitranscriptome Justin M. Thomas, Pedro J. Batista, and Jordan L. Meier ACS Chem. Biol., Just Accepted Manuscript • DOI: 10.1021/acschembio.8b00951 • Publication Date (Web): 17 Jan 2019 Downloaded from http://pubs.acs.org on January 20, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

245x109mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

211x95mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 2 of 29

Page 3 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

146x133mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

155x61mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 4 of 29

Page 5 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

76x46mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

Metabolic Regulation of the Epitranscriptome

2 3

Justin M. Thomas[a], Pedro J. Batista*[b] and Jordan L. Meier*[a]

4 5

[a]Chemical

6

[b]Laboratory

Biology Laboratory, National Cancer Institute, Frederick MD, 21702, USA.. of Cell Biology, National Cancer Institute, Bethesda MD, 20892, USA..

7 8

*Email: [email protected], [email protected]

9 10

Keywords

11

Oncometabolite: Metabolites whose abnormal accumulation causes altered signaling that can

12

lead to cellular transformation and malignancy.

13 14

Epitranscriptome: Shorthand used to refer to RNA modifications, by analogy with epigenome.

15 16

Cofactor competition: Mechanism of metabolic regulation whereby an endogenous metabolite

17

competes with substrate for enzyme active site occupancy, similar to a synthetic inhibitor.

18 19

Cofactor depletion: Mechanism of metabolic regulation whereby depletion of a critical enzyme

20

cofactor (e.g. SAM) reduces the activity of enzymes with high Michaelis constant (Km) for that

21

cofactor.

22 23

Metabolic regulation of writer localization: Mechanism of metabolic regulation whereby

24

cofactor concentrations may affect affinity and interaction of epitransciptomic writers with RNA

25

substrates.

ACS Paragon Plus Environment

Page 6 of 29

Page 7 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

26

Abstract:

27

An emergent theme in cancer biology is that dysregulated energy metabolism may directly

28

influence oncogenic gene expression. This is due to the fact that many enzymes involved in gene

29

regulation use utilize cofactors derived from primary metabolism, including acetyl-CoA, S-

30

adenosylmethionine, and 2-ketoglutarate. While this phenomenon was first studied through the

31

prism of histone and DNA modifications (the epigenome), recent work indicates metabolism can

32

also impact gene regulation by disrupting the balance of RNA post-transcriptional modifications

33

(the epitranscriptome). Here we review recent studies that explore how the metabolic regulation

34

of writers and erasers of the epitranscriptome (FTO, TET2, NAT10, MTO1, METTL16) helps

35

shape gene expression through three distinct mechanisms: cofactor inhibition, cofactor depletion,

36

and writer localization. Our brief survey underscores similarities and differences between the

37

metabolic regulation of the epigenome and epitranscriptome, and highlights fertile ground for

38

future investigation.

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

39

Introduction

40

The ability of metabolism to influence gene regulatory mechanisms in human cells is an

41

increasingly well-established phenomenon. While canonical metabolic sensors such as growth

42

factor receptors and nutrient-sensitive kinases have been known for decades, a more recent

43

finding is that metabolites may play a direct role in transcriptional regulation by influencing

44

chromatin structure in the nucleus. Early studies in this area demonstrated that disruption of

45

acetyl-CoA metabolism can inhibit histone acetylation, implying that under certain conditions

46

acetyl-CoA becomes rate-limiting for histone acetyltransferases.1-2 These findings were

47

remarkable, as they suggested the enzymatic that add (“writers”) and remove (“erasers”) of

48

histone modifications (many of which use primary metabolites as cofactors) may serve as a link

49

between environmental cues and powerful epigenetic mechanisms of gene regulation in

50

eukaryotes. This concept of ‘metabolic regulation of epigenetics,’ as well examples of its

51

relevance to physiology and disease, has been summarized excellently elsewhere.3-5

52

Five years ago in ACS Chemical Biology, we reviewed three mechanisms by which

53

metabolites had been found to directly impact the activity of enzymes involved in epigenetic

54

signaling: 1) competitive inhibition, 2) cofactor depletion, and 3) localized enzymatic effects.6

55

Intriguingly, these metabolic mechanisms are not necessarily limited to enzymes involved in DNA

56

and histone modifications but may also encompass other classes of cofactor-utilizing enzymes

57

that play a signaling role in cancer, including the writers and erasers of RNA modifications. The

58

ability of RNA to function as a direct sensor of metabolic information is well-precedented in

59

bacteria, where riboswitches drive nutrient-dependent gene expression programs.7 Related is the

60

fact that over 100 modified nucleobases exist in human coding and non-coding RNA and, just like

61

histone modifications, many of these modifications are derived from primary metabolic cofactors.

62

Similarly, just as histone modifications are associated with changes in gene transcription, RNA

63

modifications can alter gene expression via diverse mechanisms.8-10 The importance of these

64

mechanisms is highlighted by the critical roles described for RNA modification in both

ACS Paragon Plus Environment

Page 8 of 29

Page 9 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

65

development and in disease.11 In this perspective we summarize recent efforts to understand the

66

metabolic regulation of RNA modifications, also known as the epitranscriptome. Rather than

67

providing a comprehensive review of the metabolic regulation of RNA modifications, this article

68

aims to focus attention on a few contemporary examples of studies of metabolically-sensitive

69

writer and eraser enzymes as case studies for defining this phenomenon. In addition to reviewing

70

these critical studies, our brief survey highlights similarities and differences between the metabolic

71

regulation of the epigenome and epitranscriptome, and suggests future areas for further

72

exploration.

73 74

Metabolic Regulation of the Epitranscriptome by Competitive Inhibition (FTO, TET2,

75

NAT10)

76

The most disease-relevant mechanism by which metabolism has been shown to directly influence

77

protein and nucleic acid modifications is via the production of competitive metabolites. The

78

hallmark example of this was found in 2008, when large-scale glioblastoma and acute myeloid

79

leukemia genome sequencing efforts identified driver mutations in the active-sites of the primary

80

metabolic enzymes isocitrate dehydrogenase 1 and 2 (IDH1/IDH2).12-14 Mechanistic studies

81

revealed these genetic lesions cause formation of heterodimeric IDH enzymes imbued with novel

82

ability to transform 2-ketoglutarate into (R)-2-hydroxyglutarate (2-HG).15 (R)-2-HG accumulates

83

to millimolar levels in tumors and patient tissues, allowing it to competitively inhibit Fe(II)--

84

ketoglutarate-dependent dioxygenase enzymes involved in control of histone demethylation, and

85

DNA hydroxymethylation.16-17 The ability to promote malignant transformation by influencing gene

86

regulatory mechanisms has led (R)-2-HG to be termed an “oncometabolite.”

87

In contrast to their ability to trigger cancer initiation, an increasingly appreciated

88

paradoxical property of oncometabolites is their ability to exert deleterious effects on mature tumor

89

progression. This is exemplified by the finding that IDH mutant glioblastomas are associated with

90

a longer median overall survival than other variants of the disease,18-19 and that high levels of (R)-

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

91

2-HG can inhibit cell growth and confer collateral inhibitor vulnerabilities across a number of cell

92

types.20-22 However, it is relatively unknown how specific dioxygenases communicate information

93

about (R)-2-HG levels to gene expression programs involved in regulation of cell growth. Su and

94

coworkers examined the effects of (R)-2-HG on the enzyme FTO, an Fe(II)--ketoglutarate

95

dioxygenase which catalyzes the removal of N6-methyladenosine (m6A) and N6,2’-O-

96

dimethyladenosine (m6Am) from RNA.23-25 Careful analysis of gene expression data found that

97

sensitivity to (R)-2-HG correlated with high expression of FTO. This led the authors to hypothesize

98

these cell lines may be addicted to active FTO-dependent m6A demethylation, which preliminary

99

data from an earlier study had suggested.26 Consistent with this hypothesis, treatment of AML cell

100

lines with (R)-2-HG led to an increase overall m6A levels. (R)-2-HG also caused an increase in

101

the proteolytic and thermal stability of FTO, providing additional support for ligand occupancy and

102

a direct oncometabolite-FTO interaction.27-28 Transcriptome-wide mapping of m6A sites found (R)-

103

2-HG treatment increased the methylation of 5’-untranslated regions (5’-UTRs) and coding

104

sequences of many transcripts, including the master oncogenic transcription factor Myc.

105

Expression of a synthetic gene construct incorporating the Myc 5’-UTR was activated by FTO

106

overexpression and inhibited by (R)-2-HG treatment. These and additional experiments support

107

an overall model in which competitive inhibition of FTO by (R)-2-HG increases levels of m6A in

108

the Myc 5’-UTR, accelerating decay of Myc mRNA and limiting pro-growth gene expression

109

programs (Figure 1). (R)-2-HG also increased m6A -dependent decay of CEBPA, a transcription

110

factor necessary for FTO expression, providing an additional mechanism that amplifies

111

oncometabolite-dependent m6A disequilibrium.

112

These findings have several implications for the biology of IDH mutant cancers. First, they

113

suggest that disruption of the FTO/ m6A/Myc axis may be an important contributor to the improved

114

prognosis observed in cancer harboring driver mutations in IDH enzymes. Since this initial study

115

focused on examining the response of m6A to (R)-2-HG levels raised by dosing or ectopic

116

expression of IDH mutants, it will be important to determine whether m6A equilibrium is similarly

ACS Paragon Plus Environment

Page 10 of 29

Page 11 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

117

altered in AML clinical specimens harboring neomorphic IDH mutations. Furthermore, the finding

118

that (R)-2-HG antagonizes an m6A demethylation program required for optimal cell growth

119

suggests inhibition of FTO may be a viable therapeutic strategy in IDH mutant cancers. FTO has

120

recently been found to play an oncogenic role in a variety of mutationally-defined AMLs, and it is

121

tempting to speculate that the partial inhibition of FTO caused by (R)-2-HG accumulation may

122

render IDH mutant cancers especially sensitive to small molecule FTO inhibitors.29 Related to this

123

thought, the finding that after malignant transformation (R)-2-HG production actually limits cell

124

growth suggests in some contexts these cancers may be better treated by targeting collateral

125

vulnerabilities30 (such as FTO), rather than directly targeting production of (R)-2-HG itself using

126

inhibitors of mutant IDH.31-32 Future work will be required to test this hypothesis, as well as to

127

assess how m6A levels correlate with clinical response. Another question raised by these studies

128

is whether modulation of m6A may play a role in gene regulation in other settings marked by (R)-

129

2-HG accumulation. In addition to other IDH mutant cancers,33 elevated (R)-2-HG levels have

130

been observed in the absence of IDH mutation, and hpoxia has been found to increase levels of

131

(S)-2-HG.34-35 It is likely that examination of m6A and m6Am in these diverse settings will yield

132

additional insights into the metabolic regulation of RNA modifications and their role in malignancy.

133

Extending our analogy, in the metabolic regulation of epigenetic signaling, cofactor

134

competitive metabolites have been found to alter the activity of many different classes of

135

chromatin-modifying enzymes.36 Similarly, early evidence suggests these active-site dependent

136

mechanisms may also apply to several classes of RNA modifiers. The TET proteins are Fe(II)--

137

ketoglutarate-dependent enzymes that oxidize 5-methylcytidine to 5-hydroxymethylcytidine

138

(hm5C) in DNA and RNA.37-38 (R)-2-HG has been validated as a competitive inhibitor of TET2,39

139

and TET2 and IDH mutations are mutually exclusive in AML,40 suggesting they carry out

140

redundant functions. A recent study found that ectopic overexpression of neomorphic IDH1 and

141

IDH2 mutants altered the ability of TET2 to catalyze formation of hm5C in RNA.41 Coupled with

142

recent evidence that intragenic hm5C is present in mRNA and can functionally alter translation,42

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

143

these findings suggest decreased levels of hm5C in RNA may contribute to gene expression

144

changes in IDH mutant AML.

145

Projecting beyond 2-HG, many methyltransferases have been found to be sensitive to S-

146

adenosylmethionine (SAM)-competitive metabolites such as S-methylthioadenosine (MTA),

147

which accumulates in methylthioadenosine phosphorylase (MTAP)-deficient cancers.43 It is

148

plausible that elevated levels of MTA may also alter the activity of RNA methyltransferase

149

enzymes, which are responsible for >40 eukaryotic RNA modifications that have been postulated

150

to be derived from SAM. Recently an acetylated nucleobase, N4-acetylcytidine (ac4C), was found

151

to occur in human transfer RNA (tRNA) and ribosomal RNA (rRNA).44-45 Our group applied an

152

unbiased chemical proteomic approach46 to discover that NAT10, the enzyme responsible for

153

ac4C deposition, binds tightly to the metabolic feedback inhibitor CoA.47 This suggests NAT10

154

may be sensitive to the cellular acetyl-CoA/CoA ratio, which is known to be decreased by fasting

155

and bioenergetic stress.1-2 Consistent with this hypothesis, one group has reported that disruption

156

of acetyl-CoA biosynthesis reduces levels of ac4C in yeast,48 while another has found that

157

starvation of C. elegans inhibits ac4C levels in tRNA.49 Although a definitive role for ac4C in gene

158

regulation awaits determination, this modification is highly conserved and has been implicated in

159

control of tRNA stability, rRNA biogenesis, and mRNA stability.45, 50-51 Intriguingly, the activity of

160

NAT10 has also been implicated in premature aging syndromes.52 Given the well-characterized

161

links between metabolism, histone acetylation, and aging,53-54 the development of new methods

162

to study RNA acetylation55-56 may help determine whether this modification also play a role in the

163

gene regulatory response to stimuli such as caloric restriction and Warburg metabolism.

164 165

Metabolic Regulation of the Epitranscriptome by Cofactor Depletion (MTO1, GTPBP3)

166

A second way in which metabolism can influence the activity of cofactor-utilizing enzymes is via

167

cofactor depletion.6 In this mechanism, depletion of a critical metabolic cofactor (e.g. SAM)

168

reduces the activity of enzymes possess a high Michaelis constant (Km) for that cofactor.

ACS Paragon Plus Environment

Page 12 of 29

Page 13 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

169

Despite much study, it remains undetermined whether cofactor depletion alone can cause rate-

170

limiting decreases in the activity of enzymes that utilize SAM, 2-ketoglutarate, and acetyl-CoA,

171

or whether these activity changes are reliant on coincidental increases in competitive

172

metabolites such as SAH, 2-HG, and CoA (described above). However, a recent study by

173

Morscher and coworkers suggests that cofactor depletion can powerfully repress the activity of

174

RNA-modifying enzymes responsible for synthesis of the tRNA modification 5-

175

taurinomethyluridine (τm5U).57 These studies were spurred by the curious observation that

176

cellular knockout of SHMT2, but not other folate one-carbon (1C) biosynthetic enzymes,

177

enhanced extracellular acidification of colon cancer cell grown in culture. SHMT2 plays a critical

178

role in folate 1C metabolism by mediating the serine-dependent transformation of THF to

179

methylene-THF in the mitochondria.58 Hypothesizing that the observed increase in glycolysis

180

could be due to impaired mitochondrial function, Morscher et al. used a variety of functional

181

assays to determine that deletion of SHMT2 impairs cellular respiration by decreasing the

182

abundance of the complex I, IV, and V subunits of the mitochondrial electron transport chain.

183

Careful metabolic detective work and feeding studies further revealed methylene-THF as the

184

key metabolite whose depletion by SHMT2 knockout was responsible for respiratory chain

185

deficiency. After ruling out mechanisms such as mitochondrial DNA damage, the authors next

186

considered the possibility that SHMT2/methylene-THF depletion may specifically inhibit

187

translation in the mitochondria.

188

Mitochondria contain localized ribosomes and tRNAs that are responsible for the

189

translation of 13 proteins in humans, including several enzymes in complex I, IV, and V of the

190

electron transport chain.59 To test the hypothesis that methylene-THF may be required for

191

protein translation in this organelle, a mitochondria-specific ribosome profiling method was

192

developed. The deep coverage and high resolution afforded by this method revealed that

193

SHMT2 knockout decreased levels of actively translating ribosomes at mRNAs encoding

194

several subunits of the respiratory complex, in part due to a high rate of stalled ribosomes at the

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

195

aminoacyl-tRNA acceptor site of specific lysine and leucine codons (LysAAG and LeuUUG).57 No

196

stalling was observed at LysAAA and LeuUUA codons, suggesting that SHMT2 deficiency causes

197

cells difficulty reading the 3’ guanosine of certain codons. In humans, decoding of codons

198

ending in A/G in split codon boxes is facilitated by modifications of the tRNA anticodon which

199

allow non-Watson-Crick base-pairing with the codon 3’ base. Within mitochondria this is

200

accomplished by tRNA wobble base modifications derived from 1C metabolism, including 5’-

201

taurinomethyluridine (τm5U). To determine whether τm5U may be sensitive to mitochondrial

202

methylene-THF levels, the authors assessed this modification by LC-MS and found that τm5U

203

and its 2-thio derivative were depleted to undetectable levels in SHMT2 knockout cell lines. This

204

reduction was not due to reduced taurine in these cell lines and could be rescued by re-

205

expression of wild-type SHMT2 or administration of the methylene-THF precursor sarcosine.

206

Isotopic labeling experiments revealed the metabolic source of the τm5U methyl group was

207

methylene-THF rather than SAM, denoting τm5U as the first macromolecular modification

208

directly derived from folate in mammals. MTO1 is a member of the enzyme complex responsible

209

for formation of τm5U in human mitochondrial tRNAs. Consistent with the hypothesis that

210

methylene-THF can become rate-limiting for τm5U deposition, knockout of the τm5U biosynthetic

211

enzyme MTO1 increased occupancy at the same 3’-guanosine-containing codons affected by

212

SHMT2 depletion. In addition, evidence suggests disrupted folate 1C metabolism may inhibit

213

τm5U -mediated translation of respiratory chain proteins in several additional contexts, including

214

mitochondrial disorders, neonatal folate deficiency, and the therapeutic administration of

215

antifolates to treat immune disorders and cancer.57 Valuably, parallel investigations of τm5U in

216

mitochondrial disease made similar findings regarding its regulation by folate metabolism and

217

dietary taurine, providing additional perspective and validation.60-61

218

The discovery of a mitochondrial translational program regulated by 1C metabolism

219

raises several questions for the field. First, what is the molecular basis for the sensitivity of the

220

τm5U modification to folate metabolism? In this regard it is noteworthy that 5-formylcytidine

ACS Paragon Plus Environment

Page 14 of 29

Page 15 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

221

(f5C), a 1C metabolism-derived RNA modification found at the 5’-anticodon of mitochondrial

222

tRNAMet, was unaffected by SHMT2 ablation.41 This suggests that metabolic cofactor depletion

223

can selectively impact the activity of a subset of RNA-modifying enzymes, a concept similar to

224

the spectrum of metabolic sensitivity displayed by chromatin-modifying enzymes such as

225

histone acetyltransferases. In the case of τm5U and f5C, this selectivity may stem from a

226

different range of cellular metabolite concentrations attained their 1C metabolite cofactors

227

(methylene-THF and SAM), differential enzyme kinetic parameters such as cofactor Km, or

228

some combination thereof. Further biochemical and metabolomic analyses will likely be

229

informative in this regard. An additional question raised by these studies is whether metabolic

230

mechanisms regulate codon-specific translation via RNA modifications in other contexts. As the

231

role of tRNA modifications in the cellular stress response has been detailed elsewhere,10, 62-63

232

we limit ourselves here to one potentially informative example. In yeast, levels of sulfur-

233

containing amino acids modulate levels of 5-methoxycarbonylmethyl-2-thiouridine (mcm5s2U), a

234

modified nucleobase found at the 5’-anticodon (wobble base) of several cytosolic tRNAs.64 65-67

235

While nutrient-sensitive mcm5s2U deposition has not yet been reported in humans, it is known

236

that altered levels of the enzymes which synthesize this modification can alter gene expression

237

through codon-biased translation in neurodegenerative diseases as well as cancer.65, 68 In the

238

future it will be interesting to assess whether mcm5s2U is dysregulated in cancer contexts that

239

are highly dependent on cysteine,69-70 or responsive to changes in acetyl-CoA and SAM levels,

240

which also contribute to formation of this modification and contain cysteine-derived elements.71

241

Complex indirect regulation is also plausible, as exemplified by the recent finding that 1-

242

methyladenosine in tRNA regulates translation and is responsive to glucose-dependent gene

243

expression changes.72 Beyond these tRNA examples, gene translation may also be altered by

244

mRNA modifications, some of which have already established to be metabolically sensitive

245

(including m6A detailed above).73 It is likely that as novel roles for RNA modifications in gene

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

246

expression continue to be discovered, additional examples of metabolic regulation of the

247

epitranscriptome by cofactor depletion mechanism will emerge.

248 249

Metabolic Regulation of Epitranscriptome Writer Localization. (METTL16)

250

Besides globally turning the activity of their writer and eraser enzymes, metabolic stimuli can also

251

manifest surprising effects at specific genomic and transcriptomic loci.6 One of the first examples

252

of this mechanism was characterized in chromatin, where genomic localization of the SAM

253

synthetase MAT2A was found to be required for histone methylation and transcriptional

254

repression of specific genes.74 This supported a model in which MAT2A supplies a localized pool

255

of SAM to protein methyltransferases to aid transcriptional regulation. Extending this paradigm to

256

the epitranscriptome, Pendleton and coworkers recently demonstrated that metabolic

257

perturbations can reduce localized RNA methyltransferase activity at the MAT2A transcript level,

258

and in so doing elicit a protein-RNA interaction which aids maintenance of cellular SAM levels.75

259

MAT2A is a metabolic enzyme responsible for catalyzing the ATP-dependent formation of

260

SAM from methionine. Early studies had found that the abundance of the MAT2A mRNA

261

increases upon methionine depletion, presumably to allow cells to maintain SAM levels during

262

nutrient stress.76 Through a series of elegant experiments, Pendleton et al determined that

263

reduced cellular SAM triggers the splicing of an intron-retained MAT2A isoform that is normally

264

unstable. Splicing of the intron-retained isoform increases the abundance of the MAT2A mRNA,

265

and was further found to depend on a proximal hairpin in the 3’-UTR which is marked by the

266

methylated RNA nucleobase m6A. Mutational studies in a model MAT2A construct demonstrated

267

that methionine-dependent splicing an m6A site in the proximal hairpin. Additional studies found

268

that methylation did not depend on the major mRNA methyltransferase METTL3,77 but rather

269

catalyzed by a relatively uncharacterized methyltransferase, METTL16. Unexpectedly, while

270

methionine depletion and METTL16 knockdown both decreased m6A in MAT2A’s 3’-UTR, only

271

methionine depletion prompted splicing. This led the authors to propose a model where inefficient

ACS Paragon Plus Environment

Page 16 of 29

Page 17 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

272

methylation due to low SAM causes METTL16 to linger on the MAT2A transcript. This metabolite-

273

dependent increase in occupancy then promotes METTL16-driven splicing through m6A-

274

independent effects. Consistent with this, the authors found that overexpression of a METTL16

275

catalytic domain mutant, but not an RNA-binding domain mutant, could trigger splicing. Similarly,

276

using MS2-tethering to increase METTL16 binding to a synthetic MAT2A construct increased

277

splicing. Mutational analyses together with RNA immunoprecipitation were used to determine that

278

the METTL16 methyltransferase domain was required for methionine-dependent changes in

279

MAT2A mRNA occupancy, while the highly conserved VCR domain was found to be required for

280

splicing. A number of additional studies to further characterize the regulation of m6A by METTL16

281

led to the discovery that METTL16 can have a broad influence on m6A in messenger RNA through

282

a combination of direct and indirect effects, a subset of which may be mediated by MAT2A.75 In

283

another example of scientific convergence, analogous findings regarding the feedback regulation

284

of MAT2A mRNA stability were also made by Shima and coworkers.78

285

The discovery of a functional metabolite-dependent protein-RNA interaction provides

286

fodder for future investigation of several areas. First, these studies suggest that the absence of

287

an enzyme cofactor can cause RNA modification writers such as METTL16 to functionally act as

288

readers, binding to unmodified substrates constitutively, and triggering non-catalytic functions

289

such as nucleation of protein-protein interactions. This recalls studies that have suggested

290

chromatin-modifying histone deacetylases can also act as functional readers.79 The ability of this

291

mechanism to influence activity would also be expected to be highly dependent on kinetic

292

mechanism of these enzymes; for example, many acetyltransferases are known exhibit ordered

293

binding of acetyl-CoA followed by peptide substrate,80 while many methyltransferases are known

294

to interact with substrate first, or show no preference for ordered binding.81 In the future it will be

295

interesting to understand the biochemical basis for high occupancy RNA recognition by METTL16,

296

as well as whether this paradigm applies to other classes of writers and erasers of nucleic acid

297

and protein modifications. Of note, many writers and erasers of modifications are multifunctional

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

298

enzymes with important non-catalytic activities, including the major mRNA methyltransferase

299

METTL3 as well as the histone acetyltransferase EP300.82-83 Another question pertains to how

300

prevalent the splicing-based stabilization of nuclear intron isoforms is as a homeostatic

301

mechanism. A recent report found this mechanism regulates OGT, a glycosyltransferase that

302

utilizes the glucose-derived metabolite UDP-O-GlcNAc as a metabolic precursor.84 The ability of

303

cells to rapidly increase OGT levels may explain observations that glucose-derived histone

304

acetylation is sensitive nutrient stress but protein O-GlcNAc-ylation is not.2, 85 This intimates a

305

larger point, which is the ability of homeostatic mechanisms to cause unexpected effects on

306

metabolically-derived protein and nucleic acid modifications. For example, based on the

307

mechanism above MAT2A knockdown would be expected to have a more substantial effect on

308

SAM levels than methionine deprivation. Thus, studying the metabolic regulation of an RNA

309

modification under these two different conditions may lead to dramatically different conclusions.

310

This suggests that care should be taken to directly measure metabolite levels, as well as probe

311

for the existence of homeostatic feedback mechanisms based on writer, reader, eraser, or

312

biosynthetic enzyme abundance in future studies of metabolically-regulated signaling. While

313

homeostatic feedback adds a layer of complexity onto the paradigm of metabolic control, the

314

studies above provide a model for unraveling these mechanisms and also powerfully illustrate the

315

ability of such studies to yield new insights into gene regulation.

316 317

Conclusions

318

Here we have reviewed several recent examples of how metabolic mechanisms can influence the

319

posttranscriptional chemical modification of RNA, also known as the epitranscriptome. In stem

320

cells populations, metabolism is rewired in order to maintain a balance between pluripotency and

321

differentiation, a strategy co-opted by several cancers. Compared to proteins and DNA, RNAs

322

undergoes a larger spectrum of enzyme-catalyzed modifications, many of which are sourced from

323

a more complex pool of metabolites. Exemplifying this, the mitochondrial wobble base

ACS Paragon Plus Environment

Page 18 of 29

Page 19 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

324

modification τm5U is derived from methylene-THF and taurine, while the cytosolic modification

325

mcm5s2U derives from cysteine, SAM, and acetyl-CoA (Figure 4). This suggests the potential for

326

RNA modifications to have a more complex relationship with metabolism than protein

327

modifications. This attribute is potentially compounded by the sequential establishment of many

328

RNA modifications, affording additional opportunity for interplay. Furthermore, the participation of

329

multiple classes of RNA molecules, all targets of writers and erasers, in the processes leading to

330

protein production further contributes to a complex interaction between the metabolic environment

331

in the cell and posttranscriptional regulation of gene expression. Another difference between

332

epigenetic and epitranscriptomic regulation is the relative time-scale of these processes. While

333

chromatin modifications can be copied during replication and inherited by daughter cells, RNA

334

modifications will be lost upon turnover. The limited lifetime of cellular RNAs may alleviate the

335

necessity for eraser enzymes for many of these modifications, with the consequence that the

336

direct influence of metabolism on RNA modifications may manifest primarily through writer

337

enzymes. Conversely, while changes in the equilibrium of chromatin modifications require

338

extended time periods to be established, the relatively rapid turnover of RNA provides a

339

mechanism for more rapid response to environmental stimuli, as exemplified by the FTO/Myc

340

axis. Defining how these kinetic differences influence the scope of RNA modification-dependent

341

regulatory responses, how metabolism may affect RNA stability and turnover, and whether rapid

342

changes in RNA modifications can help facilitate long-term changes to the epigenome will be

343

important to future understanding of this phenomenon. Finally, it is important to emphasize that

344

we have highlighted only a small number of examples where metabolism has been found to

345

directly influence the writers and erasers of RNA modifications. Additional metabolic mechanisms

346

of epitranscriptomic regulation have been proposed86 and characterized,87 and many more novel

347

examples no doubt await discovery. Future efforts to unravel these mechanisms will benefit from

348

the development and application of new tools to study metabolic regulation. Methods for rapid

349

metabolite induction and depletion will be important to study fast biological responses prior to

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

350

engagement of pleiotropic effects.88-89 Improved approaches for mapping mRNA modifications

351

may enable their effects on gene regulation to be assessed with greater resolution.90-91 New tools

352

for the high-throughput analysis of metabolite-protein interactions involved in the regulation of

353

writer and eraser enzymes will be useful to identify new pathways susceptible to this

354

phenomenon.92-94 Further efforts in this area have the potential to provide substantial insights into

355

the signaling role of metabolites and the role they play in the powerful gene regulatory

356

mechanisms that underlie fundamental biology, medicine, and human disease.

357 358

Acknowledgements.

359

We sincerely apologize to the many authors whose work could not be included due to space

360

limitations. This work was supported by National Institutes of Health, National Cancer Institute,

361

Center for Cancer Research (ZIABC011766 and ZIABC011488).

362

ACS Paragon Plus Environment

Page 20 of 29

Page 21 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

363

ACS Chemical Biology

Figure Legends.

364 365

Figure 1. Metabolic regulation of the epitranscriptome by competitive inhibition. Mutant IDH1

366

produces high levels of (R)-2-hydroxyglutarate which competitively inhibits FTO, a Fe(II)/2-

367

ketoglutarate-dependent dioxygenase that erases m6A in mRNA. Increased m6A levels in mRNAs

368

encoding oncogenes such as Myc reduce transcript stability and impede cell growth.

369 370

Figure 2. Metabolic regulation of the epitranscriptome by cofactor depletion. Inhibition of SHMT2

371

depletes 5,10-methylene-THF, which serves as the 1C donor for the enzymatic formation of 5-

372

taurinomethyluridine (τm5U). This reduces τm5U in a subset of mitochondrial tRNAs and causes

373

decreased translation of electron transport chain proteins.

374 375

Figure 3. Metabolic regulation of epitranscriptome writer localization. Under conditions of nutrient

376

abundance (left), METTL16 catalyzes methylation of an mRNA encoding the SAM biosynthetic

377

enzyme MAT2A, limiting its splicing and reducing MAT2A mRNA stability. When SAM levels are

378

low (right), METTL16 inefficiently methylates the MAT2A mRNA but forms a stable interaction.

379

This triggers splicing and mRNA stabilization that allows more MAT2A to be produced and for

380

cellular SAM homeostasis to be maintained.

381 382

Figure 4. Diverse metabolic sources of RNA modifications. m6A = N6-methyladenosine, mcm5s2U

383

= 5-methoxycarbonylmethyl-2-thiouridine, ac4C = N4-acetylcytidine.

384

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

385 386

References

387 388

1. Cai, L.; Sutter, B. M.; Li, B.; Tu, B. P., Acetyl-CoA induces cell growth and proliferation by promoting the acetylation of histones at growth genes. Mol Cell 2011, 42, 426-437.

389 390 391

2. Wellen, K. E.; Hatzivassiliou, G.; Sachdeva, U. M.; Bui, T. V.; Cross, J. R.; Thompson, C. B., ATP-citrate lyase links cellular metabolism to histone acetylation. Science 2009, 324, 1076-1080.

392

3.

393 394

4. Kinnaird, A.; Zhao, S.; Wellen, K. E.; Michelakis, E. D., Metabolic control of epigenetics in cancer. Nat Rev Cancer 2016, 16, 694-707.

395 396

5. Schvartzman, J. M.; Thompson, C. B.; Finley, L. W. S., Metabolic regulation of chromatin modifications and gene expression. J Cell Biol 2018, 217, 2247-2259.

397 398

6. Meier, J. L., Metabolic Mechanisms of Epigenetic Regulation. Acs Chemical Biology 2013, 8, 2607-2621.

399 400

7. McCown, P. J.; Corbino, K. A.; Stav, S.; Sherlock, M. E.; Breaker, R. R., Riboswitch diversity and distribution. RNA 2017, 23, 995-1011.

401 402 403

8. Sloan, K. E.; Warda, A. S.; Sharma, S.; Entian, K. D.; Lafontaine, D. L. J.; Bohnsack, M. T., Tuning the ribosome: The influence of rRNA modification on eukaryotic ribosome biogenesis and function. RNA Biol 2017, 14, 1138-1152.

404 405

9. Zhao, B. S.; Roundtree, I. A.; He, C., Post-transcriptional gene regulation by mRNA modifications. Nat Rev Mol Cell Biol 2017, 18, 31-42.

406 407

10. Pan, T., Adaptive translation as a mechanism of stress response and adaptation. Annu Rev Genet 2013, 47, 121-137.

408 409

11. Frye, M.; Harada, B. T.; Behm, M.; He, C., RNA modifications modulate gene expression during development. Science 2018, 361, 1346-1349.

410 411 412 413 414 415

12. Parsons, D. W.; Jones, S.; Zhang, X.; Lin, J. C.; Leary, R. J.; Angenendt, P.; Mankoo, P.; Carter, H.; Siu, I. M.; Gallia, G. L.; Olivi, A.; McLendon, R.; Rasheed, B. A.; Keir, S.; Nikolskaya, T.; Nikolsky, Y.; Busam, D. A.; Tekleab, H.; Diaz, L. A., Jr.; Hartigan, J.; Smith, D. R.; Strausberg, R. L.; Marie, S. K.; Shinjo, S. M.; Yan, H.; Riggins, G. J.; Bigner, D. D.; Karchin, R.; Papadopoulos, N.; Parmigiani, G.; Vogelstein, B.; Velculescu, V. E.; Kinzler, K. W., An integrated genomic analysis of human glioblastoma multiforme. Science 2008, 321, 1807-1812.

416 417 418 419 420 421 422

13. Mardis, E. R.; Ding, L.; Dooling, D. J.; Larson, D. E.; McLellan, M. D.; Chen, K.; Koboldt, D. C.; Fulton, R. S.; Delehaunty, K. D.; McGrath, S. D.; Fulton, L. A.; Locke, D. P.; Magrini, V. J.; Abbott, R. M.; Vickery, T. L.; Reed, J. S.; Robinson, J. S.; Wylie, T.; Smith, S. M.; Carmichael, L.; Eldred, J. M.; Harris, C. C.; Walker, J.; Peck, J. B.; Du, F.; Dukes, A. F.; Sanderson, G. E.; Brummett, A. M.; Clark, E.; McMichael, J. F.; Meyer, R. J.; Schindler, J. K.; Pohl, C. S.; Wallis, J. W.; Shi, X.; Lin, L.; Schmidt, H.; Tang, Y.; Haipek, C.; Wiechert, M. E.; Ivy, J. V.; Kalicki, J.; Elliott, G.; Ries, R. E.; Payton, J. E.; Westervelt, P.; Tomasson, M. H.; Watson,

Lu, C.; Thompson, C. B., Metabolic regulation of epigenetics. Cell Metab 2012, 16, 9-17.

ACS Paragon Plus Environment

Page 22 of 29

Page 23 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

423 424 425

M. A.; Baty, J.; Heath, S.; Shannon, W. D.; Nagarajan, R.; Link, D. C.; Walter, M. J.; Graubert, T. A.; DiPersio, J. F.; Wilson, R. K.; Ley, T. J., Recurring mutations found by sequencing an acute myeloid leukemia genome. N Engl J Med 2009, 361, 1058-1066.

426 427

14. Balss, J.; Meyer, J.; Mueller, W.; Korshunov, A.; Hartmann, C.; von Deimling, A., Analysis of the IDH1 codon 132 mutation in brain tumors. Acta Neuropathol 2008, 116, 597-602.

428 429 430 431 432

15. Dang, L.; White, D. W.; Gross, S.; Bennett, B. D.; Bittinger, M. A.; Driggers, E. M.; Fantin, V. R.; Jang, H. G.; Jin, S.; Keenan, M. C.; Marks, K. M.; Prins, R. M.; Ward, P. S.; Yen, K. E.; Liau, L. M.; Rabinowitz, J. D.; Cantley, L. C.; Thompson, C. B.; Vander Heiden, M. G.; Su, S. M., Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature 2009, 462, 739744.

433 434

16. L, M. G.; Boulay, K.; Topisirovic, I.; Huot, M. E.; Mallette, F. A., Oncogenic Activities of IDH1/2 Mutations: From Epigenetics to Cellular Signaling. Trends Cell Biol 2017, 27, 738-752.

435 436 437 438

17. Lu, C.; Ward, P. S.; Kapoor, G. S.; Rohle, D.; Turcan, S.; Abdel-Wahab, O.; Edwards, C. R.; Khanin, R.; Figueroa, M. E.; Melnick, A.; Wellen, K. E.; O'Rourke, D. M.; Berger, S. L.; Chan, T. A.; Levine, R. L.; Mellinghoff, I. K.; Thompson, C. B., IDH mutation impairs histone demethylation and results in a block to cell differentiation. Nature 2012, 483, 474-478.

439 440 441 442 443 444

18. Patel, J. P.; Gonen, M.; Figueroa, M. E.; Fernandez, H.; Sun, Z.; Racevskis, J.; Van Vlierberghe, P.; Dolgalev, I.; Thomas, S.; Aminova, O.; Huberman, K.; Cheng, J.; Viale, A.; Socci, N. D.; Heguy, A.; Cherry, A.; Vance, G.; Higgins, R. R.; Ketterling, R. P.; Gallagher, R. E.; Litzow, M.; van den Brink, M. R.; Lazarus, H. M.; Rowe, J. M.; Luger, S.; Ferrando, A.; Paietta, E.; Tallman, M. S.; Melnick, A.; Abdel-Wahab, O.; Levine, R. L., Prognostic relevance of integrated genetic profiling in acute myeloid leukemia. N Engl J Med 2012, 366, 1079-1089.

445 446 447 448

19. Sanson, M.; Marie, Y.; Paris, S.; Idbaih, A.; Laffaire, J.; Ducray, F.; El Hallani, S.; Boisselier, B.; Mokhtari, K.; Hoang-Xuan, K.; Delattre, J. Y., Isocitrate dehydrogenase 1 codon 132 mutation is an important prognostic biomarker in gliomas. J Clin Oncol 2009, 27, 41504154.

449 450 451

20. Karpel-Massler, G.; Ishida, C. T.; Bianchetti, E.; Zhang, Y.; Shu, C.; Tsujiuchi, T.; Banu, M. A.; Garcia, F.; Roth, K. A.; Bruce, J. N.; Canoll, P.; Siegelin, M. D., Induction of synthetic lethality in IDH1-mutated gliomas through inhibition of Bcl-xL. Nat Commun 2017, 8, 1067.

452 453 454

21. Chan, S. M.; Thomas, D.; Corces-Zimmerman, M. R.; Xavy, S.; Rastogi, S.; Hong, W. J.; Zhao, F.; Medeiros, B. C.; Tyvoll, D. A.; Majeti, R., Isocitrate dehydrogenase 1 and 2 mutations induce BCL-2 dependence in acute myeloid leukemia. Nat Med 2015, 21, 178-184.

455 456 457 458 459 460

22. Sulkowski, P. L.; Corso, C. D.; Robinson, N. D.; Scanlon, S. E.; Purshouse, K. R.; Bai, H.; Liu, Y.; Sundaram, R. K.; Hegan, D. C.; Fons, N. R.; Breuer, G. A.; Song, Y.; Mishra-Gorur, K.; De Feyter, H. M.; de Graaf, R. A.; Surovtseva, Y. V.; Kachman, M.; Halene, S.; Gunel, M.; Glazer, P. M.; Bindra, R. S., 2-Hydroxyglutarate produced by neomorphic IDH mutations suppresses homologous recombination and induces PARP inhibitor sensitivity. Sci Transl Med 2017, 9.

461 462 463

23. Su, R.; Dong, L.; Li, C.; Nachtergaele, S.; Wunderlich, M.; Qing, Y.; Deng, X.; Wang, Y.; Weng, X.; Hu, C.; Yu, M.; Skibbe, J.; Dai, Q.; Zou, D.; Wu, T.; Yu, K.; Weng, H.; Huang, H.; Ferchen, K.; Qin, X.; Zhang, B.; Qi, J.; Sasaki, A. T.; Plas, D. R.; Bradner, J. E.; Wei, M.;

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

464 465

Marcucci, G.; Jiang, X.; Mulloy, J. C.; Jin, J.; He, C.; Chen, J., R-2HG Exhibits Anti-tumor Activity by Targeting FTO/m(6)A/MYC/CEBPA Signaling. Cell 2018, 172, 90-105 e123.

466 467 468

24. Jia, G.; Fu, Y.; Zhao, X.; Dai, Q.; Zheng, G.; Yang, Y.; Yi, C.; Lindahl, T.; Pan, T.; Yang, Y. G.; He, C., N6-methyladenosine in nuclear RNA is a major substrate of the obesityassociated FTO. Nat Chem Biol 2011, 7, 885-887.

469 470 471 472

25. Mauer, J.; Luo, X.; Blanjoie, A.; Jiao, X.; Grozhik, A. V.; Patil, D. P.; Linder, B.; Pickering, B. F.; Vasseur, J. J.; Chen, Q.; Gross, S. S.; Elemento, O.; Debart, F.; Kiledjian, M.; Jaffrey, S. R., Reversible methylation of m(6)Am in the 5' cap controls mRNA stability. Nature 2017, 541, 371-375.

473 474 475

26. Elkashef, S. M.; Lin, A. P.; Myers, J.; Sill, H.; Jiang, D.; Dahia, P. L. M.; Aguiar, R. C. T., IDH Mutation, Competitive Inhibition of FTO, and RNA Methylation. Cancer Cell 2017, 31, 619620.

476 477 478 479

27. Lomenick, B.; Hao, R.; Jonai, N.; Chin, R. M.; Aghajan, M.; Warburton, S.; Wang, J.; Wu, R. P.; Gomez, F.; Loo, J. A.; Wohlschlegel, J. A.; Vondriska, T. M.; Pelletier, J.; Herschman, H. R.; Clardy, J.; Clarke, C. F.; Huang, J., Target identification using drug affinity responsive target stability (DARTS). Proc Natl Acad Sci U S A 2009, 106, 21984-21989.

480 481 482

28. Martinez Molina, D.; Jafari, R.; Ignatushchenko, M.; Seki, T.; Larsson, E. A.; Dan, C.; Sreekumar, L.; Cao, Y.; Nordlund, P., Monitoring drug target engagement in cells and tissues using the cellular thermal shift assay. Science 2013, 341, 84-87.

483 484 485 486 487

29. Li, Z.; Weng, H.; Su, R.; Weng, X.; Zuo, Z.; Li, C.; Huang, H.; Nachtergaele, S.; Dong, L.; Hu, C.; Qin, X.; Tang, L.; Wang, Y.; Hong, G. M.; Huang, H.; Wang, X.; Chen, P.; Gurbuxani, S.; Arnovitz, S.; Li, Y.; Li, S.; Strong, J.; Neilly, M. B.; Larson, R. A.; Jiang, X.; Zhang, P.; Jin, J.; He, C.; Chen, J., FTO Plays an Oncogenic Role in Acute Myeloid Leukemia as a N(6)Methyladenosine RNA Demethylase. Cancer Cell 2017, 31, 127-141.

488 489 490 491 492 493 494

30. McBrayer, S. K.; Mayers, J. R.; DiNatale, G. J.; Shi, D. D.; Khanal, J.; Chakraborty, A. A.; Sarosiek, K. A.; Briggs, K. J.; Robbins, A. K.; Sewastianik, T.; Shareef, S. J.; Olenchock, B. A.; Parker, S. J.; Tateishi, K.; Spinelli, J. B.; Islam, M.; Haigis, M. C.; Looper, R. E.; Ligon, K. L.; Bernstein, B. E.; Carrasco, R. D.; Cahill, D. P.; Asara, J. M.; Metallo, C. M.; Yennawar, N. H.; Vander Heiden, M. G.; Kaelin, W. G., Jr., Transaminase Inhibition by 2-Hydroxyglutarate Impairs Glutamate Biosynthesis and Redox Homeostasis in Glioma. Cell 2018, 175, 101-116 e125.

495 496 497 498 499 500 501

31. DiNardo, C. D.; Stein, E. M.; de Botton, S.; Roboz, G. J.; Altman, J. K.; Mims, A. S.; Swords, R.; Collins, R. H.; Mannis, G. N.; Pollyea, D. A.; Donnellan, W.; Fathi, A. T.; Pigneux, A.; Erba, H. P.; Prince, G. T.; Stein, A. S.; Uy, G. L.; Foran, J. M.; Traer, E.; Stuart, R. K.; Arellano, M. L.; Slack, J. L.; Sekeres, M. A.; Willekens, C.; Choe, S.; Wang, H.; Zhang, V.; Yen, K. E.; Kapsalis, S. M.; Yang, H.; Dai, D.; Fan, B.; Goldwasser, M.; Liu, H.; Agresta, S.; Wu, B.; Attar, E. C.; Tallman, M. S.; Stone, R. M.; Kantarjian, H. M., Durable Remissions with Ivosidenib in IDH1-Mutated Relapsed or Refractory AML. N Engl J Med 2018, 378, 2386-2398.

502 503 504 505

32. Wang, F.; Travins, J.; DeLaBarre, B.; Penard-Lacronique, V.; Schalm, S.; Hansen, E.; Straley, K.; Kernytsky, A.; Liu, W.; Gliser, C.; Yang, H.; Gross, S.; Artin, E.; Saada, V.; Mylonas, E.; Quivoron, C.; Popovici-Muller, J.; Saunders, J. O.; Salituro, F. G.; Yan, S.; Murray, S.; Wei, W.; Gao, Y.; Dang, L.; Dorsch, M.; Agresta, S.; Schenkein, D. P.; Biller, S. A.; Su, S. M.; de

ACS Paragon Plus Environment

Page 24 of 29

Page 25 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

506 507

Botton, S.; Yen, K. E., Targeted inhibition of mutant IDH2 in leukemia cells induces cellular differentiation. Science 2013, 340, 622-626.

508 509

33. Yang, H.; Ye, D.; Guan, K. L.; Xiong, Y., IDH1 and IDH2 mutations in tumorigenesis: mechanistic insights and clinical perspectives. Clin Cancer Res 2012, 18, 5562-5571.

510 511 512

34. Intlekofer, A. M.; Dematteo, R. G.; Venneti, S.; Finley, L. W.; Lu, C.; Judkins, A. R.; Rustenburg, A. S.; Grinaway, P. B.; Chodera, J. D.; Cross, J. R.; Thompson, C. B., Hypoxia Induces Production of L-2-Hydroxyglutarate. Cell Metab 2015, 22, 304-311.

513 514 515

35. Oldham, W. M.; Clish, C. B.; Yang, Y.; Loscalzo, J., Hypoxia-Mediated Increases in L-2hydroxyglutarate Coordinate the Metabolic Response to Reductive Stress. Cell Metab 2015, 22, 291-303.

516 517

36. Meier, J. L., Metabolic mechanisms of epigenetic regulation. ACS Chem Biol 2013, 8, 2607-2621.

518 519 520

37. Fu, L.; Guerrero, C. R.; Zhong, N.; Amato, N. J.; Liu, Y.; Liu, S.; Cai, Q.; Ji, D.; Jin, S. G.; Niedernhofer, L. J.; Pfeifer, G. P.; Xu, G. L.; Wang, Y., Tet-mediated formation of 5hydroxymethylcytosine in RNA. J Am Chem Soc 2014, 136, 11582-11585.

521 522 523

38. Tahiliani, M.; Koh, K. P.; Shen, Y.; Pastor, W. A.; Bandukwala, H.; Brudno, Y.; Agarwal, S.; Iyer, L. M.; Liu, D. R.; Aravind, L.; Rao, A., Conversion of 5-methylcytosine to 5hydroxymethylcytosine in mammalian DNA by MLL partner TET1. Science 2009, 324, 930-935.

524 525 526

39. Losman, J. A.; Looper, R. E.; Koivunen, P.; Lee, S.; Schneider, R. K.; McMahon, C.; Cowley, G. S.; Root, D. E.; Ebert, B. L.; Kaelin, W. G., Jr., (R)-2-hydroxyglutarate is sufficient to promote leukemogenesis and its effects are reversible. Science 2013, 339, 1621-1625.

527 528 529 530 531 532

40. Figueroa, M. E.; Abdel-Wahab, O.; Lu, C.; Ward, P. S.; Patel, J.; Shih, A.; Li, Y.; Bhagwat, N.; Vasanthakumar, A.; Fernandez, H. F.; Tallman, M. S.; Sun, Z.; Wolniak, K.; Peeters, J. K.; Liu, W.; Choe, S. E.; Fantin, V. R.; Paietta, E.; Lowenberg, B.; Licht, J. D.; Godley, L. A.; Delwel, R.; Valk, P. J.; Thompson, C. B.; Levine, R. L.; Melnick, A., Leukemic IDH1 and IDH2 mutations result in a hypermethylation phenotype, disrupt TET2 function, and impair hematopoietic differentiation. Cancer Cell 2010, 18, 553-567.

533 534

41. Xu, Q.; Wang, K.; Wang, L.; Zhu, Y.; Zhou, G.; Xie, D.; Yang, Q., IDH1/2 Mutants Inhibit TET-Promoted Oxidation of RNA 5mC to 5hmC. PLoS One 2016, 11, e0161261.

535 536 537 538 539

42. Delatte, B.; Wang, F.; Ngoc, L. V.; Collignon, E.; Bonvin, E.; Deplus, R.; Calonne, E.; Hassabi, B.; Putmans, P.; Awe, S.; Wetzel, C.; Kreher, J.; Soin, R.; Creppe, C.; Limbach, P. A.; Gueydan, C.; Kruys, V.; Brehm, A.; Minakhina, S.; Defrance, M.; Steward, R.; Fuks, F., RNA biochemistry. Transcriptome-wide distribution and function of RNA hydroxymethylcytosine. Science 2016, 351, 282-285.

540 541 542 543 544

43. Kryukov, G. V.; Wilson, F. H.; Ruth, J. R.; Paulk, J.; Tsherniak, A.; Marlow, S. E.; Vazquez, F.; Weir, B. A.; Fitzgerald, M. E.; Tanaka, M.; Bielski, C. M.; Scott, J. M.; Dennis, C.; Cowley, G. S.; Boehm, J. S.; Root, D. E.; Golub, T. R.; Clish, C. B.; Bradner, J. E.; Hahn, W. C.; Garraway, L. A., MTAP deletion confers enhanced dependency on the PRMT5 arginine methyltransferase in cancer cells. Science 2016, 351, 1214-1218.

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

545 546 547

44. Ito, S.; Horikawa, S.; Suzuki, T.; Kawauchi, H.; Tanaka, Y.; Suzuki, T.; Suzuki, T., Human NAT10 is an ATP-dependent RNA acetyltransferase responsible for N4-acetylcytidine formation in 18 S ribosomal RNA (rRNA). J Biol Chem 2014, 289, 35724-35730.

548 549 550 551

45. Sharma, S.; Langhendries, J. L.; Watzinger, P.; Kotter, P.; Entian, K. D.; Lafontaine, D. L., Yeast Kre33 and human NAT10 are conserved 18S rRNA cytosine acetyltransferases that modify tRNAs assisted by the adaptor Tan1/THUMPD1. Nucleic Acids Res 2015, 43, 22422258.

552 553

46. Montgomery, D. C.; Sorum, A. W.; Meier, J. L., Defining the orphan functions of lysine acetyltransferases. ACS Chem Biol 2015, 10, 85-94.

554 555 556

47. Montgomery, D. C.; Garlick, J. M.; Kulkarni, R. A.; Kennedy, S.; Allali-Hassani, A.; Kuo, Y. M.; Andrews, A. J.; Wu, H.; Vedadi, M.; Meier, J. L., Global Profiling of Acetyltransferase Feedback Regulation. J Am Chem Soc 2016, 138, 6388-6391.

557 558 559

48. Ito, S.; Akamatsu, Y.; Noma, A.; Kimura, S.; Miyauchi, K.; Ikeuchi, Y.; Suzuki, T.; Suzuki, T., A single acetylation of 18 S rRNA is essential for biogenesis of the small ribosomal subunit in Saccharomyces cerevisiae. J Biol Chem 2014, 289, 26201-26212.

560 561 562

49. van Delft, P.; Akay, A.; Huber, S. M.; Bueschl, C.; Rudolph, K. L. M.; Di Domenico, T.; Schuhmacher, R.; Miska, E. A.; Balasubramanian, S., The Profile and Dynamics of RNA Modifications in Animals. Chembiochem 2017, 18, 979-984.

563 564 565

50. Dewe, J. M.; Whipple, J. M.; Chernyakov, I.; Jaramillo, L. N.; Phizicky, E. M., The yeast rapid tRNA decay pathway competes with elongation factor 1A for substrate tRNAs and acts on tRNAs lacking one or more of several modifications. RNA 2012, 18, 1886-1896.

566 567 568 569

51. Arango, D.; Sturgill, D.; Alhusaini, N.; Dillman, A. A.; Sweet, T. J.; Hanson, G.; Hosogane, M.; Sinclair, W. R.; Nanan, K. K.; Mandler, M. D.; Fox, S. D.; Zengeya, T. T.; Andresson, T.; Meier, J. L.; Coller, J.; Oberdoerffer, S., Acetylation of Cytidine in mRNA Promotes Translation Efficiency. Cell 2018.

570 571

52. Larrieu, D.; Britton, S.; Demir, M.; Rodriguez, R.; Jackson, S. P., Chemical inhibition of NAT10 corrects defects of laminopathic cells. Science 2014, 344, 527-532.

572 573

53. Ali, I.; Conrad, R. J.; Verdin, E.; Ott, M., Lysine Acetylation Goes Global: From Epigenetics to Metabolism and Therapeutics. Chem Rev 2018, 118, 1216-1252.

574 575

54. Shimazu, T.; Hirschey, M. D.; Huang, J. Y.; Ho, L. T.; Verdin, E., Acetate metabolism and aging: An emerging connection. Mech Ageing Dev 2010, 131, 511-516.

576 577 578

55. Sinclair, W. R.; Arango, D.; Shrimp, J. H.; Zengeya, T. T.; Thomas, J. M.; Montgomery, D. C.; Fox, S. D.; Andresson, T.; Oberdoerffer, S.; Meier, J. L., Profiling Cytidine Acetylation with Specific Affinity and Reactivity. ACS Chem Biol 2017, 12, 2922-2926.

579 580 581 582

56. Thomas, J. M.; Briney, C. A.; Nance, K. D.; Lopez, J. E.; Thorpe, A. L.; Fox, S. D.; Bortolin-Cavaille, M. L.; Sas-Chen, A.; Arango, D.; Oberdoerffer, S.; Cavaille, J.; Andresson, T.; Meier, J. L., A Chemical Signature for Cytidine Acetylation in RNA. J Am Chem Soc 2018, 140, 12667-12670.

ACS Paragon Plus Environment

Page 26 of 29

Page 27 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

583 584 585

57. Morscher, R. J.; Ducker, G. S.; Li, S. H.; Mayer, J. A.; Gitai, Z.; Sperl, W.; Rabinowitz, J. D., Mitochondrial translation requires folate-dependent tRNA methylation. Nature 2018, 554, 128-132.

586 587

58. Tibbetts, A. S.; Appling, D. R., Compartmentalization of Mammalian folate-mediated one-carbon metabolism. Annu Rev Nutr 2010, 30, 57-81.

588 589

59. D'Souza, A. R.; Minczuk, M., Mitochondrial transcription and translation: overview. Essays Biochem 2018, 62, 309-320.

590 591 592 593 594

60. Asano, K.; Suzuki, T.; Saito, A.; Wei, F. Y.; Ikeuchi, Y.; Numata, T.; Tanaka, R.; Yamane, Y.; Yamamoto, T.; Goto, T.; Kishita, Y.; Murayama, K.; Ohtake, A.; Okazaki, Y.; Tomizawa, K.; Sakaguchi, Y.; Suzuki, T., Metabolic and chemical regulation of tRNA modification associated with taurine deficiency and human disease. Nucleic Acids Res 2018, 46, 1565-1583.

595 596

61. Lucas, S.; Chen, G.; Aras, S.; Wang, J., Serine catabolism is essential to maintain mitochondrial respiration in mammalian cells. Life Sci Alliance 2018, 1, e201800036.

597 598 599

62. Endres, L.; Dedon, P. C.; Begley, T. J., Codon-biased translation can be regulated by wobble-base tRNA modification systems during cellular stress responses. RNA Biol 2015, 12, 603-614.

600 601

63. Gu, C.; Begley, T. J.; Dedon, P. C., tRNA modifications regulate translation during cellular stress. FEBS Lett 2014, 588, 4287-4296.

602 603 604

64. Laxman, S.; Sutter, B. M.; Wu, X.; Kumar, S.; Guo, X.; Trudgian, D. C.; Mirzaei, H.; Tu, B. P., Sulfur amino acids regulate translational capacity and metabolic homeostasis through modulation of tRNA thiolation. Cell 2013, 154, 416-429.

605 606 607 608 609 610

65. Rapino, F.; Delaunay, S.; Rambow, F.; Zhou, Z.; Tharun, L.; De Tullio, P.; Sin, O.; Shostak, K.; Schmitz, S.; Piepers, J.; Ghesquiere, B.; Karim, L.; Charloteaux, B.; Jamart, D.; Florin, A.; Lambert, C.; Rorive, A.; Jerusalem, G.; Leucci, E.; Dewaele, M.; Vooijs, M.; Leidel, S. A.; Georges, M.; Voz, M.; Peers, B.; Buttner, R.; Marine, J. C.; Chariot, A.; Close, P., Codonspecific translation reprogramming promotes resistance to targeted therapy. Nature 2018, 558, 605-609.

611 612

66. Johansson, M. J. O.; Xu, F.; Bystrom, A. S., Elongator-a tRNA modifying complex that promotes efficient translational decoding. Biochim Biophys Acta 2018, 1861, 401-408.

613 614

67. Kojic, M.; Wainwright, B., The Many Faces of Elongator in Neurodevelopment and Disease. Front Mol Neurosci 2016, 9, 115.

615 616

68. Landgraf, B. J.; McCarthy, E. L.; Booker, S. J., Radical S-Adenosylmethionine Enzymes in Human Health and Disease. Annu Rev Biochem 2016, 85, 485-514.

617 618 619 620

69. Briggs, K. J.; Koivunen, P.; Cao, S.; Backus, K. M.; Olenchock, B. A.; Patel, H.; Zhang, Q.; Signoretti, S.; Gerfen, G. J.; Richardson, A. L.; Witkiewicz, A. K.; Cravatt, B. F.; Clardy, J.; Kaelin, W. G., Jr., Paracrine Induction of HIF by Glutamate in Breast Cancer: EglN1 Senses Cysteine. Cell 2016, 166, 126-139.

ACS Paragon Plus Environment

ACS Chemical Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

621 622 623

70. Muir, A.; Danai, L. V.; Gui, D. Y.; Waingarten, C. Y.; Lewis, C. A.; Vander Heiden, M. G., Environmental cystine drives glutamine anaplerosis and sensitizes cancer cells to glutaminase inhibition. Elife 2017, 6.

624 625

71. Selvadurai, K.; Wang, P.; Seimetz, J.; Huang, R. H., Archaeal Elp3 catalyzes tRNA wobble uridine modification at C5 via a radical mechanism. Nat Chem Biol 2014, 10, 810-812.

626 627 628

72. Liu, F.; Clark, W.; Luo, G.; Wang, X.; Fu, Y.; Wei, J.; Wang, X.; Hao, Z.; Dai, Q.; Zheng, G.; Ma, H.; Han, D.; Evans, M.; Klungland, A.; Pan, T.; He, C., ALKBH1-Mediated tRNA Demethylation Regulates Translation. Cell 2016, 167, 1897.

629 630 631 632

73. Choi, J.; Ieong, K. W.; Demirci, H.; Chen, J.; Petrov, A.; Prabhakar, A.; O'Leary, S. E.; Dominissini, D.; Rechavi, G.; Soltis, S. M.; Ehrenberg, M.; Puglisi, J. D., N(6)-methyladenosine in mRNA disrupts tRNA selection and translation-elongation dynamics. Nat Struct Mol Biol 2016, 23, 110-115.

633 634 635

74. Katoh, Y.; Ikura, T.; Hoshikawa, Y.; Tashiro, S.; Ito, T.; Ohta, M.; Kera, Y.; Noda, T.; Igarashi, K., Methionine adenosyltransferase II serves as a transcriptional corepressor of Maf oncoprotein. Mol Cell 2011, 41, 554-566.

636 637 638

75. Pendleton, K. E.; Chen, B.; Liu, K.; Hunter, O. V.; Xie, Y.; Tu, B. P.; Conrad, N. K., The U6 snRNA m(6)A Methyltransferase METTL16 Regulates SAM Synthetase Intron Retention. Cell 2017, 169, 824-835 e814.

639 640 641 642

76. Martinez-Chantar, M. L.; Latasa, M. U.; Varela-Rey, M.; Lu, S. C.; Garcia-Trevijano, E. R.; Mato, J. M.; Avila, M. A., L-methionine availability regulates expression of the methionine adenosyltransferase 2A gene in human hepatocarcinoma cells: role of S-adenosylmethionine. J Biol Chem 2003, 278, 19885-19890.

643 644 645

77. Liu, J.; Yue, Y.; Han, D.; Wang, X.; Fu, Y.; Zhang, L.; Jia, G.; Yu, M.; Lu, Z.; Deng, X.; Dai, Q.; Chen, W.; He, C., A METTL3-METTL14 complex mediates mammalian nuclear RNA N6-adenosine methylation. Nat Chem Biol 2014, 10, 93-95.

646 647 648 649

78. Shima, H.; Matsumoto, M.; Ishigami, Y.; Ebina, M.; Muto, A.; Sato, Y.; Kumagai, S.; Ochiai, K.; Suzuki, T.; Igarashi, K., S-Adenosylmethionine Synthesis Is Regulated by Selective N(6)-Adenosine Methylation and mRNA Degradation Involving METTL16 and YTHDC1. Cell Rep 2017, 21, 3354-3363.

650 651 652

79. Bradner, J. E.; West, N.; Grachan, M. L.; Greenberg, E. F.; Haggarty, S. J.; Warnow, T.; Mazitschek, R., Chemical phylogenetics of histone deacetylases. Nat Chem Biol 2010, 6, 238243.

653 654

80. Tanner, K. G.; Langer, M. R.; Denu, J. M., Kinetic mechanism of human histone acetyltransferase P/CAF. Biochemistry 2000, 39, 11961-11969.

655 656 657

81. Patnaik, D.; Chin, H. G.; Esteve, P. O.; Benner, J.; Jacobsen, S. E.; Pradhan, S., Substrate specificity and kinetic mechanism of mammalian G9a histone H3 methyltransferase. J Biol Chem 2004, 279, 53248-53258.

658 659

82. Lin, S.; Choe, J.; Du, P.; Triboulet, R.; Gregory, R. I., The m(6)A Methyltransferase METTL3 Promotes Translation in Human Cancer Cells. Mol Cell 2016, 62, 335-345.

ACS Paragon Plus Environment

Page 28 of 29

Page 29 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

660 661

83. Bedford, D. C.; Brindle, P. K., Is histone acetylation the most important physiological function for CBP and p300? Aging (Albany NY) 2012, 4, 247-255.

662 663 664

84. Park, S. K.; Zhou, X.; Pendleton, K. E.; Hunter, O. V.; Kohler, J. J.; O'Donnell, K. A.; Conrad, N. K., A Conserved Splicing Silencer Dynamically Regulates O-GlcNAc Transferase Intron Retention and O-GlcNAc Homeostasis. Cell Rep 2017, 20, 1088-1099.

665 666 667

85. Taylor, R. P.; Parker, G. J.; Hazel, M. W.; Soesanto, Y.; Fuller, W.; Yazzie, M. J.; McClain, D. A., Glucose deprivation stimulates O-GlcNAc modification of proteins through upregulation of O-linked N-acetylglucosaminyltransferase. J Biol Chem 2008, 283, 6050-6057.

668 669 670

86. Bird, J. G.; Zhang, Y.; Tian, Y.; Panova, N.; Barvik, I.; Greene, L.; Liu, M.; Buckley, B.; Krasny, L.; Lee, J. K.; Kaplan, C. D.; Ebright, R. H.; Nickels, B. E., The mechanism of RNA 5' capping with NAD+, NADH and desphospho-CoA. Nature 2016, 535, 444-447.

671 672

87. Kiledjian, M., Eukaryotic RNA 5'-End NAD(+) Capping and DeNADding. Trends Cell Biol 2018, 28, 454-464.

673 674 675 676

88. Nabet, B.; Roberts, J. M.; Buckley, D. L.; Paulk, J.; Dastjerdi, S.; Yang, A.; Leggett, A. L.; Erb, M. A.; Lawlor, M. A.; Souza, A.; Scott, T. G.; Vittori, S.; Perry, J. A.; Qi, J.; Winter, G. E.; Wong, K. K.; Gray, N. S.; Bradner, J. E., The dTAG system for immediate and target-specific protein degradation. Nat Chem Biol 2018, 14, 431-441.

677 678 679

89. Holland, A. J.; Fachinetti, D.; Han, J. S.; Cleveland, D. W., Inducible, reversible system for the rapid and complete degradation of proteins in mammalian cells. Proc Natl Acad Sci U S A 2012, 109, E3350-3357.

680 681 682

90. Safra, M.; Sas-Chen, A.; Nir, R.; Winkler, R.; Nachshon, A.; Bar-Yaacov, D.; Erlacher, M.; Rossmanith, W.; Stern-Ginossar, N.; Schwartz, S., The m1A landscape on cytosolic and mitochondrial mRNA at single-base resolution. Nature 2017, 551, 251-255.

683 684 685

91. Shu, X.; Dai, Q.; Wu, T.; Bothwell, I. R.; Yue, Y.; Zhang, Z.; Cao, J.; Fei, Q.; Luo, M.; He, C.; Liu, J., N(6)-Allyladenosine: A New Small Molecule for RNA Labeling Identified by Mutation Assay. J Am Chem Soc 2017, 139, 17213-17216.

686 687 688 689 690

92. Kulkarni, R. A.; Worth, A. J.; Zengeya, T. T.; Shrimp, J. H.; Garlick, J. M.; Roberts, A. M.; Montgomery, D. C.; Sourbier, C.; Gibbs, B. K.; Mesaros, C.; Tsai, Y. C.; Das, S.; Chan, K. C.; Zhou, M.; Andresson, T.; Weissman, A. M.; Linehan, W. M.; Blair, I. A.; Snyder, N. W.; Meier, J. L., Discovering Targets of Non-enzymatic Acylation by Thioester Reactivity Profiling. Cell Chem Biol 2017, 24, 231-242.

691 692 693 694 695

93. Joberty, G.; Boesche, M.; Brown, J. A.; Eberhard, D.; Garton, N. S.; Humphreys, P. G.; Mathieson, T.; Muelbaier, M.; Ramsden, N. G.; Reader, V.; Rueger, A.; Sheppard, R. J.; Westaway, S. M.; Bantscheff, M.; Lee, K.; Wilson, D. M.; Prinjha, R. K.; Drewes, G., Interrogating the Druggability of the 2-Oxoglutarate-Dependent Dioxygenase Target Class by Chemical Proteomics. ACS Chem Biol 2016, 11, 2002-2010.

696 697 698 699

94. Hashimoto, M.; Girardi, E.; Eichner, R.; Superti-Furga, G., Detection of Chemical Engagement of Solute Carrier Proteins by a Cellular Thermal Shift Assay. ACS Chem Biol 2018, 13, 1480-1486.

ACS Paragon Plus Environment