Accepted Manuscript Pyrimidine Metabolism: Dynamic and Versatile Pathways in Pathogens and Cellular Development Manuel F. Garavito, Heidy Y. Narváez-Ortiz, Barbara H. Zimmermann PII:

S1673-8527(15)00076-4

DOI:

10.1016/j.jgg.2015.04.004

Reference:

JGG 363

To appear in:

Journal of Genetics and Genomics

Received Date: 21 November 2014 Revised Date:

13 April 2015

Accepted Date: 14 April 2015

Please cite this article as: Garavito, M.F., Narváez-Ortiz, H.Y., Zimmermann, B.H., Pyrimidine Metabolism: Dynamic and Versatile Pathways in Pathogens and Cellular Development, Journal of Genetics and Genomics (2015), doi: 10.1016/j.jgg.2015.04.004. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT

Pyrimidine Metabolism: Dynamic and Versatile Pathways in Pathogens and Cellular

SC

RI PT

Development

M AN U

Manuel F. Garavito1, Heidy Y. Narváez-Ortiz1 and Barbara H. Zimmermann*

Departamento de Ciencias Biologicas, Universidad de los Andes, Carrera 1, No. 18A-10,

TE D

Bogotá, Colombia

*Corresponding author. Tel.: 571 339 4949 ext 3374; fax: 571 613 0222. [email protected]

authors contributed equally to this work.

AC C

‡These

EP

Email

Abbreviations: CAD, multifunctional protein that initiates and regulates de novo pyrimidine biosynthesis; DHODase, dihydroorotate dehydrogenase; hpi, hours postinfection; UMP synthase, uridine 5′-monophosphate synthase; UMP, uridine 5′-monophosphate; UTP, uridine 5′triphosphate; UPRTase, uracil phosphoribosyltransferase; UPase, uridine phosphorylase.

1

ACCEPTED MANUSCRIPT

1

Abstract

2

The importance of pyrimidines lies in the fact that they are structural components of a broad

3

spectrum of key molecules that participate in diverse cellular functions, such as synthesis of

4

DNA, RNA, lipids, and carbohydrates.

5

involved in the synthesis, degradation, salvage and interconversion and transport of these

6

molecules.

7

metabolism changes under a variety of conditions including, when possible, those studies based

8

on techniques of genomics, transcriptomics, proteomics, and metabolomics. First, we briefly

9

look at the dynamics of pyrimidine metabolism during nonpathogenic cellular events. We then

10

focus on changes that pathogen infections cause in the pyrimidine metabolism of their host.

11

Next, we discuss the effects of antimetabolites and inhibitors, and finally we consider the

12

consequences of genetic manipulations, such as knock-downs, knock-outs, and knock-ins, of

13

pyrimidine enzymes on pyrimidine metabolism in the cell.

14

Keywords:

15

dehydrogenase, UMP synthase.

16

(1) Introduction.

17

Pyrimidines are essential biomolecules. They are structural components of nucleotides, nucleic

18

acids, vitamins, pterins and folates, each fulfilling critical roles in the cell (Cox, 1968). The

19

metabolism of pyrimidines has been broadly studied in many organisms. Their critical

20

participation in processes such as RNA and DNA synthesis, formation of UDP-sugars for

21

glycosylation of proteins and lipids, and formation of CDP-activated precursors for membrane

22

phospholipids, makes them attractive for study.

RI PT

Pyrimidine metabolism encompasses all enzymes

M AN U

SC

In this review, we discuss recent publications that document how pyrimidine

AC C

EP

TE D

pyrimidine metabolism, pathogens, CAD, dihydroorotase, dihydroorotate

23

Pyrimidine derivatives have been used as anticancer, antimicrobial, antiviral, antibacterial,

24

antiprotozoal, and antifungal agents. Pharmacological applications of pyrimidine-related

25

molecules are even more diverse, ranging from antipyretics, antihypertensives, antihistaminics,

26

anti-inflammatories, analgesics, antidiabetics, antiallergics, antioxidants, anticonvulsants, to

27

central nervous system depressants (Jain et al., 2006; Chauhan and Kumar, 2013; Sharma et al.,

28

2014). The study of pyrimidine metabolism is important for better understanding of both the

2

ACCEPTED MANUSCRIPT

29

biochemistry of the use of these compounds, as well as the metabolism of drugs derived from

30

pyrimidines. Fig. 1 shows most pathways of pyrimidine metabolism. There is a consensus on what

32

constitutes the de novo and degradation pathways, however, the subdivision of pyrimidine

33

metabolism reactions into salvage or interconversion pathways is controversial. We have chosen

34

to group the reactions according to whether they involve nucleobases, nucleosides, nucleosides

35

mono-, di- or triphosphates (nucleotides).

RI PT

31

The de novo pathway is initiated with bicarbonate, glutamine and two molecules of ATP.

37

This pathway involves six enzymatic steps conserved in all species, although they have different

38

architecture depending on the organism (for detailed information see Nara et al., 2000). In most

39

prokaryotes and some simple eukaryotes, the enzymes of the de novo pathway are encoded by

40

separate genes, whereas in mammals they are encoded by only three genes, carbamoylphosphate

41

synthetase–aspartate transcarbamoylase–dihydroorotase (CAD), dihydroorotate dehydrogenase

42

(DHODase) and uridine 5′-monophosphate synthase (UMP synthase), (Jones, 1980). The

43

trifunctional enzyme CAD converts the small precursors into dihydroorotic acid that is

44

transformed into orotate by DHODase. Subsequently, the bifunctional enzyme UMP synthase

45

leads to the synthesis of uridine 5′-monophosphate (UMP). UMP could be considered one of the

46

hub molecules in pyrimidine metabolism because it is the precursor of other pyrimidine

47

nucleotides.

TE D

M AN U

SC

36

The nucleoside di- and triphosphates are formed by nucleoside kinases, and nucleoside

49

diphosphates can be converted into the deoxynucleotides by the ribonucleotide reductase

50

complex. At the bottom of Fig. 1, we show reactions involving nucleotides that are necessary for

51

forming larger biomolecules such as nucleic acids, lipids and carbohydrates (for detailed

52

information see Huang and Graves, 2003).

AC C

EP

48

53

Membrane transporters also play important roles in pyrimidine metabolism as shown at the

54

top of Fig. 1. In humans, transporters are needed for importing amino acids, nucleobases and

55

nucleosides. The solute carrier (SLC) transporter family 17 proteins permit the exchange of

56

amino acids essential for pyrimidine biosynthesis such as aspartate and glutamate, and

57

nucleotides (Sreedharan et al., 2010). Equilibrative nucleoside transporters (ENTs) are

58

responsible for transport of hydrophilic nucleosides. Concentrative nucleoside transporters

3

ACCEPTED MANUSCRIPT

59

(CNT) transport nucleobases across membranes, and transporters from the ABC transporter

60

family import nucleotides and nucleosides into the cell (Sahoo et al., 2014). Pyrimidines can be degraded to simple metabolites, which become sources of carbon and

62

nitrogen. While in some organisms such as Escherichia coli, there are multiple catabolic

63

pathways, nucleosides are catabolized mainly by two pathways, one oxidative the other reductive

64

(Loh et al., 2006). The oxidative pathway (not shown in Fig.1) has been reported only in some

65

bacteria and produces urea as a primary product, and malonic acid from uridine/cytidine, or

66

methylmalonic acid from thymine, as secondary products (Patel and West, 1987). The reductive

67

pathway (Fig. 1, right), which is present in some archaea, bacteria and eukaryotes, consists of

68

three catalytic steps to produce CO2, NH3 and ß-alanine (ß-Ala) or ß-aminoisobutyric acid (ß-

69

AIB). The nucleosides, uridine and thymidine can be reductively catabolized when they are

70

transformed to nucleobases by the action of the nucleoside phosphorylase enzymes. However,

71

cytidine must be deaminated to uridine by cytidine deaminase before entering the catabolic route

72

(Vogels and van der Drift, 1976). It has been proposed that the pyrimidine catabolic pathways

73

are important not only to maintain appropriate pyrimidine pools in cells, but also to produce ß-

74

alanine, a precursor of diverse biomolecules, for example, pantothenic acid, vitamin B5 in plants

75

and bacteria (Webb et al., 2004; Katahira and Ashihara, 2006) and a neurotransmitter in

76

mammals (Tiedje et al., 2010).

TE D

M AN U

SC

RI PT

61

Different organisms contain different subsets of enzymes in the pyrimidine pathways

78

briefly described above. Table S1 indicates which genes of pyrimidine metabolic proteins are

79

present in the organisms mentioned in this review. The presence or absence of specific

80

pyrimidine enzymes in different species probably evolved depending on the availability and

81

accessibility of nutrients in the niche in which each organism developed. It is precisely these

82

marked differences between species that makes pyrimidine metabolism a potential drug target.

83

For example, protozoan parasites of great medical importance such as Plasmodium, Toxoplasma

84

and Cryptosporidium show a reduction in the numbers of genes involved in pyrimidine

85

metabolism. Plasmodium parasites are unable to obtain pyrimidines by the salvage pathway and

86

depend completely on the de novo pathway (Fig. 2), whereas in Cryptosporidium the opposite

87

occurs (Hyde, 2007).

AC C

EP

77

88

Our goal in this review is to assemble recent advances pertinent to an overview of

89

pyrimidine metabolism. The demand for pyrimidines varies with the overall metabolic activity of

4

ACCEPTED MANUSCRIPT

the cell. It has long been known that actively dividing cells require higher levels of pyrimidines

91

than quiescent cells, thus, cancer cells are observed to overexpress pyrimidine metabolic

92

enzymes (Evans and Guy, 2004; Hu et al., 2013). Consideration of this last area is beyond the

93

scope of this review, and instead we will focus primarily, but not exclusively, on the changes in

94

pyrimidine metabolism during the life cycles of some pathogens, including those parasitic stages

95

that occur within host organism.

96

RI PT

90

Some of the first insights into the roles of pathways were gained by measuring enzyme

97

activities and by observing the effects of inhibitors.

98

permitted monitoring expression levels of enzymes in these pathways. The newest contributions

99

to the area of pyrimidine metabolism have been through metabolomics.

SC

More recently, transcriptomics has

For a very few

organisms, data from all of these areas are available. The integration of genomic, transcriptomic,

101

proteomic, and metabolomic data will allow for a better understanding of the dynamics of

102

pyrimidine metabolism.

M AN U

100

A commonly-used approach that highlights the importance of specific enzymes in a

104

pathway is to interfere with their normal function by using inhibitors. Such studies are critical

105

for identifying potential drug targets for pathogens. A more recent strategy that delineates the

106

importance of pathway enzymes is to construct genetic knock-outs, knock-ins, or knock-downs.

107

Although genetically altered organisms in many cases may have limited relevance for medical

108

applications, a study of their metabolism demonstrates the direct effects of increased or

109

decreased enzyme activities that are not plagued by unintended secondary effects or poor cell

110

permeability associated with inhibitory compounds.

EP

TE D

103

111

(2) Changes in pyrimidine metabolism during cellular development.

113

Using serum starvation to synchronize a baby hamster kidney cell line and incorporation of

AC C

112

114

14

115

pyrimidine biosynthesis in the S phase of the cell cycle was ~ 2-fold higher than in other phases

116

of the cycle (Sigoillot et al., 2003). The increase in biosynthesis could be accounted for by

117

allosteric effects on the carbamoylphosphate synthetase (CPSase) activity of CAD, which

118

included removal of feedback inhibition by UTP and increased sensitivity to the activator

119

phosphoribosyl pyrophosphate (PRPP), and was mediated by phosphorylation of CAD at Thr-

120

456 by MAP kinase (MAPK).

C-labeled bicarbonate into UTP and CTP, Sigoillot and co-workers found that the rate of

As cell cultures became confluent, Thr-456 became

5

ACCEPTED MANUSCRIPT

dephosphorylated, and cAMP-dependent kinase A (PKA) phosphorylated Ser-1406, leading to

122

decreased PRPP sensitivity, and consequently decreased CAD CPSase activity. The authors

123

measured intracellular nucleotide concentrations and observed that the levels remained

124

approximately constant during the cell cycle, and were about half the concentrations found in

125

exponentially growing cells. This led to the conclusion that newly synthesized nucleotides are

126

used as fast as they are produced (Sigoillot et al., 2003), suggesting that the pyrimidine pools in

127

these cells are maintained at subsistence levels. Two reports in 2013 (Ben-Sahra et al., 2013;

128

Robitaille et al., 2013) indicated that in addition to the allosteric regulation via the MAPK

129

cascade (Erk1/2) signaling pathway, CAD is also subject to allosteric regulation via the

130

mTORC1 signaling network which controls S6 kinase (S6K) phosphorylation of CAD at residue

131

Ser-1859.

SC

The yeast Candida albicans is an opportunistic pathogen that causes fungal infections in

M AN U

132

RI PT

121

133

immunocompromised patients.

134

electrophoresis to measure the proteins present in stationary and exponential growth phases of

135

the pathogen. The stationary phase is probably most relevant to infections because growth is

136

constrained by the nutrient-limited environment of the host cell. Five enzymes of pyrimidine

137

metabolism were among those identified.

138

orotidine 5′-monophosphate decarboxylase (ODCase), and uracil phosphoribosyltransferase

139

(UPRTase) showed increased levels in the exponential phase, and cytidine triphosphate synthase

140

(CTP synthase) showed higher levels in the stationary phase, while dihydroorotase (DHOase)

141

levels were the same in both phases (Kusch et al., 2008). Keeping in mind the caveat that higher

142

protein levels do not prove that there is increased flux through a pathway, one could explain the

143

observed changes in protein levels as follows. An increase in levels of proteins involved in

144

synthesis or recycling of nucleotides might be expected because it could allow the yeast to

145

maintain its supply of nucleotides permitting the rapid growth rate of the exponential phase. The

146

observation of the same protein levels for the DHOase in both phases may simply mean that it

147

does not play a rate-limiting role in the de novo pathway in this organism. The reason for higher

148

requirements of CTP synthase in the stationary phase is not clear, and may suggest that this

149

protein plays additional roles in the physiology of that phase. This is not surprising because it

150

has been recently documented that CTP synthase could also have a structural role in the cell (for

151

detailed information see Chen et al., 2011; Liu, 2011; Barry et al., 2014; Carcamo et al., 2014).

Kusch and co-workers (Kusch et al., 2008) used 2D-gel

AC C

EP

TE D

Orotate phosphoribosyltransferase (OPRTase),

6

ACCEPTED MANUSCRIPT

Cell death can be viewed as the expected final event in cell development. For this reason,

153

understanding the changes in the metabolic environment that lead to and occur during

154

programmed cell death (PCD) is important. As might be expected, changes are observed in

155

pyrimidine metabolism during PCD. Stasolla and co-workers (Stasolla et al., 2004) incubated

156

plant cell cultures with radiolabeled uracil, uridine and thymidine, and measured their

157

incorporation into nucleotides and nucleic acids. Degradation was also monitored. PCD was

158

induced by exposing the cultures to agents generating NO and H2O2. An early event, occurring

159

before cytological hallmarks of PCD were evident, was the reduced incorporation of uracil into

160

nucleotides and nucleic acids. Later events of PCD included reduced incorporation of uridine,

161

and continued reduced incorporation of uracil, and increased degradation of both compounds. In

162

contrast, thymidine incorporation increased (Stasolla et al., 2004). The changes in the balance of

163

salvage and degradation of pyrimidines were proposed to constitute a metabolic switch that

164

eventually commits the cell to PCD (Stasolla et al., 2004).

M AN U

SC

RI PT

152

165 166

(3) Changes in pyrimidine metabolism during pathogen infection of host cells. To ensure their own survival, many pathogens redirect the metabolism of the cells they

168

invade. In this section we will consider the changes in host pyrimidine metabolism observed in

169

two host/pathogen systems, first, fibroblasts infected by human cytomegalovirus (HCMV), and

170

second, red blood cells infected by Plasmodium falciparum. Another case we will consider is

171

that of Phytophthora infestans infecting Solanum tuberosum (potato), which is an example of a

172

pathogen modifying expression of its own pyrimidine metabolic enzymes during its life cycle.

173

This last point is also illustrated in Fig. 2, which assembles RNA Sequencing (RNA-seq) data of

174

pyrimidine metabolic enzymes for seven pathogens.

EP

TE D

167

Munger and co-workers measured metabolites during the infection of HCMV in quiescent

176

fibroblasts (Munger et al., 2006). The time course of the infection was marked by the following

177

events: viral gene expression starting at 2 hours postinfection (hpi), viral DNA replication

178

between 24 – 30 hpi, peak virus yielding between 72 – 96 hpi, and host cell death. Changes in

179

metabolites of infected cells relative to mock-infected cells were observed for the de novo

180

pyrimidine biosynthetic pathway. Carbamoyl phosphate, the product of the first step catalyzed by

181

CAD, began decreasing at 72 hpi, and at 96 hpi it was ~ 4-fold lower than mock-infected cells

182

(see Fig. 4 in Munger et al., 2006). The largest concentration change occurred with carbamoyl

AC C

175

7

ACCEPTED MANUSCRIPT

183

aspartate, the product of the second CAD reaction, which began to increase at 24 hpi, and

184

achieved levels ~ 25-fold higher than mock-infected cells at 80 hpi (see Fig. 4 in Munger et al.,

185

2006).

186

aforementioned measurements were performed with infected, serum-starved, confluent cells,

187

nevertheless, the authors found that these metabolite levels were still substantially higher than

188

those of actively growing fibroblasts.

189

significantly during infection (Munger et al., 2006). The accumulation of carbamoyl aspartate

190

might be explained as follows: the DHOase reaction is known to be reversible (Christopherson

191

and Jones, 1979), thus one could speculate that a decoupling of CAD reactions from the reaction

192

of the fourth enzyme, DHODase, could be caused by a change in localization of CAD, for

193

example from the cytosol to the nucleus (Angeletti and Engler, 1998; Sigoillot et al., 2003), and

194

might lead to reversal of the DHOase reaction. In a later publication from the same laboratory, a

195

more complete survey of metabolites was made in cells infected by HCMV and also cells

196

infected by herpes simplex virus type 1 (HSV-1) (Vastag et al., 2011). The authors concluded

197

that the two viruses employ different strategies to promote viral replication. HCMV drives the

198

host cell from quiescence toward G1/S, and the production of nucleotides, at the same time

199

preventing cellular DNA replication, while HSV-1 encodes necessary enzymes, for example

200

thymidine kinase, dUTPase, and uracil-DNA glycosylase, and increases flux through central

201

carbon metabolism to drive nucleotide synthesis (Vastag et al., 2011).

The

RI PT

UMP levels decreased slightly, while UTP, CTP, and TTP levels increased.

TE D

M AN U

SC

Interestingly, purine pool levels did not change

The first study of the metabolome in the red blood cell during the 48 h infection cycle of P.

203

falciparum revealed interesting changes in pyrimidine levels (Olszewski et al., 2009).

204

Intracellular concentrations of UMP and CMP, increased to peak concentrations of ~ 28-fold and

205

~ 24-fold, respectively, relative to uninfected cell at 32 – 48 h (see Table S1 in Olszewski et al.,

206

2009). Intracellular concentrations of the de novo synthesis intermediates dihydroorotate and

207

orotate followed a similar periodic fluctuation during the cycle, but with much lower increases (~

208

4-fold and ~ 2-fold, respectively) (see Table S1 in Olszewski et al., 2009). Concentrations in

209

extracellular media were also measured, and showed secretion of ~ 5-fold more dihydroorotate

210

from infected cells compared to uninfected cells, peaking at 32 h (see Table S2 in Olszewski et

211

al., 2009). A more dramatic increase was observed for deoxyuridine, whose secretion steadily

212

increased starting at 24 h and reached ~ 24-fold at 48 h compared to uninfected cells (see Table

213

S2 in Olszewski et al., 2009). To better understand the fluctuations in metabolites, the authors

AC C

EP

202

8

ACCEPTED MANUSCRIPT

214

measured mRNA transcript abundance during infection using DNA microarrays. Surprisingly,

215

they found a significant negative correlation between the expression levels of the enzyme

216

DHOase and the levels of its product dihydroorotate, that is, enzyme expression levels were

217

higher during initial stages of infection, when its product concentrations were low. Later in

218

infection, enzyme expression decreased, but product increased.

219

apparently contradictory data may lie in analysis of the DHOase mRNA half-life (Shock et al.,

220

2007). The biological database for the genus Plasmodium (PlasmoDB) shows that its stability is

221

low during the first 24 h, accordingly, mRNA expression must be high to ensure the

222

establishment of infection, however, after schizogony (24 h), DHOase mRNA stability is

223

enhanced by ~ 8-fold. Thus, the parasite regulates not only mRNA expression but also the

224

mRNA stability to ensure the availability of nucleotides.

SC

RI PT

The explanation for these

The life cycle of the plant pathogen P. infestans is hemibiotrophic. In the initial stages of

226

infection, the biotrophic phase, this oomycete proliferates rapidly without causing visible

227

symptoms. The later stages, known as the necrotrophic phase, lead to the death of the plant cell.

228

This change of life style is reflected in large shifts in gene expression. A quantitative reverse

229

transcription-PCR (qRT-PCR) study of relative expression profiles showed up-regulation of the

230

de novo synthetic enzymes DHOase and UMP synthase, accompanied by a concomitant down-

231

regulation of the salvage enzyme UPRTase during the biotrophic phase (García-Bayona et al.,

232

2014). During the necrotrophic phase, the relative expression showed an inverse behavior,

233

apparently relying more on salvage than on synthesis. The authors state that the changes in

234

expression could be related to higher levels of uracil, uridine, cytosine, or cytidine that may be

235

available from the dying plant cell, or could even be a strategy of the cell leading to an

236

imbalance of the nucleotide pools and host cell death phase (García-Bayona et al., 2014).

EP

TE D

M AN U

225

In Fig. 2, we have assembled transcription information for the human pathogens

238

Toxoplasma gondii (toxoplasmosis), P. falciparum (malaria), Cryptosporidium parvum

239

(cryptosporidiosis), Trypanosoma brucei (African trypanosomiasis), and Cryptococcus

240

neoformans (meningitis), and for the plant pathogens Colletotrichum higginsianum (anthracnose

241

disease) and Phytophthora ramorum (sudden oak death). Parasitic pathogens obtain metabolic

242

resources from the host cells they infect, and thus do not code for a full complement of

243

pyrimidine metabolic machinery in their genomes. Genes not present are indicated by the color

244

gray in Fig. 2. Notably, none of these pathogens have genes for pyrimidine degradation. On the

AC C

237

9

ACCEPTED MANUSCRIPT

far left side of the heat map, we have included data comparing mouse cells infected with T.

246

gondii to uninfected mouse cells. Interestingly, infection with T. gondii results in a dramatic

247

increase in uridine phosphorylase (UPase), and dihydroopyrimidine dehydrogenase in the mouse

248

cells. Comparing transcriptomes of different species using RNA-seq is challenging. One critical

249

aspect is that analyses are based on sequence conservation, yet in many cases there is a lack of

250

high-quality annotation of orthologous genes. Another difficulty in constructing this type of

251

figure was in choosing which two stages or conditions of the organism to compare, and these

252

choices were also limited by the availability of data. Additionally, it should be remembered that

253

transcriptional up- and down-regulation is just one of many mechanisms used to regulate enzyme

254

activities. Bearing these limitations in mind, what is striking in Fig. 2, is the variability in

255

transcription patterns observed for these organisms, which no doubt reflect the specialized niches

256

that they occupy.

M AN U

SC

RI PT

245

257 258

(4) Effects of antimetabolites and inhibitors on pyrimidine metabolism. An extensive study of the effect of halogenated pyrimidines was undertaken on the

260

bloodstream form of the kinetoplastid parasite T. brucei using metabolomics (Ali et al., 2013).

261

Transport assays revealed that the only pyrimidine transporter in this form of the parasite was a

262

high affinity uracil transporter, also capable of transporting orotate, uridine, and deoxyuridine,

263

but with lower affinities. The transporter exhibited 5-fold less affinity for 5-fluorouracil (5-FU)

264

than uracil, and the former was salvaged by the parasite’s UPRTase, phosphorylated, and

265

incorporated into RNA. Since no incorporation into DNA was observed, 5-FU was not a

266

substrate for parasite ribonucleotide reductase, although it is a substrate for human

267

ribonucleotide reductase. Incorporation into the UDP-activated hexose and hexosamines was

268

observed, although no differences in content of variant surface glycoprotein was found for

269

treated parasites, suggesting that 5-FU did not interfere with the parasite’s carbohydrate

270

metabolism. 5-fluoroorotate (5-FOA) was metabolized in a similar fashion to 5-FU.

271

Unexpectedly, fluoro-N-carbamoyl-L-aspartate was also detected in 5-FOA-treated parasites.

272

The authors suggested that this partial reversal of synthesis may have been caused by increased

273

levels of orotate, which they attributed to the inhibition of OPRTase by 5-FOA. They noted that

274

both compounds caused growth arrest, but did not kill the parasites quickly, and suggested that

275

their mechanism of action were probably multifactorial, and due in part to incorporation of

AC C

EP

TE D

259

10

ACCEPTED MANUSCRIPT

fluorinated nucleotides into RNA. Parasites treated with 5-fluoro-2′-deoxyuridine produced low

277

amounts of 5F-dUMP with thymidine kinase, however, no F-dUDP or F-dUTP could be

278

detected, and no incorporation into nucleic acids was observed. Treated parasites exhibited

279

elevated levels of dUMP, but had close to normal levels of thymidine nucleotides, apparently due

280

to salvage of exogenous thymidine present in the medium, supporting the idea that the effect of

281

5-fluoro-2′-deoxyuridine was likely to be caused by inhibition of thymidylate synthase.

282

The effect of increased levels of a pyrimidine intermediate may have different consequences

283

depending on cell type and conditions. Examples of this are the divergent effects of an inhibitor

284

on multiple myeloma cells (Bardeleben et al., 2013) and on hepatocellular carcinoma cells (Jung

285

et al., 2012).

286

carboxamide-1- β -riboside (AICAr), which becomes converted to AICA ribotide 5′‐

287

monophosphate (ZMP), a mimic of AMP, was found to induce apoptosis, however, this was not

288

associated with AMP-activated kinase activation or mTORC1 inhibition. A metabolomic screen

289

of AICAr-treated cells showed a ~ 26-fold increase in orotate and a ~ 0.4-fold decrease in UMP

290

that was accompanied by a 1.2 – 1.9-fold increase in purine bases and nucleotides (see Table 1 in

291

Bardeleben et al., 2013). Additionally, a ~ 3-fold decrease in PRPP levels was detected by an

292

enzymatic assay. The treated cells were found to be under replicative stress as indicated by the

293

phosphorylation of the histone H2A.X, which replaces conventional H2A in a subset of

294

nucleosomes. The authors suggested that the orotate accumulation was due to a decrease in

295

UMP synthase activity caused by limiting PRPP levels, and found that the treated cells could be

296

rescued by exogenous uridine, cytidine, and thymidine. Application of N-(phosphonacetyl)-L-

297

aspartate (PALA), a specific transition state analog inhibitor of aspartate transcarbamoylase

298

(ATCase), was also found to result in H2A.X phosphorylation, and could be rescued in a similar

299

fashion by uridine and cytidine. The authors concluded that apoptosis in AICAr-treated cells

300

was caused by pyrimidine starvation (Bardeleben et al., 2013). In contrast, in hepatocarcinoma

301

cells, exogenous application of orotate to serum-starved cells induced proliferation by increasing

302

the number of viable cells and decreasing apoptosis as mediated by mTORC1 activation (Jung et

303

al., 2012). The fate of the orotate after it was internalized by the cells was not determined (Jung

304

et al., 2012). These researchers used AICAr to activate AMP-activated protein kinase (AMPK)

305

producing negative regulation of mTOR.

306

hepatocarcinoma cells with orotate and AICAr resulted in suppression of proliferation (Jung et

SC

RI PT

276

AC C

EP

TE D

M AN U

In multiple myeloma cell cultures, the application of 5-aminoimidazole-4-

Simultaneous treatment of the starved

11

ACCEPTED MANUSCRIPT

307

al., 2012). In these two studies, AICAr appeared to affect the mTOR pathway differently,

308

however, in both cases the presence of AICAr combined with high levels of orotate, either

309

exogenous or endogenous, led to decrease in proliferation or apoptosis. In human hepatoma cell lines, inhibition of UMP synthase was shown to have potent

311

effects on cholesterol and lipid metabolism (Poirier et al., 2014). The unmethylable cytosine

312

analog 5-azacytidine (5-AzaC) is used as an epigenetic drug which permits cells to re-express

313

genes silenced by hypermethylation. Poirer and co-workers found that 5-AzaC had methylation-

314

independent effects that included decreasing cellular UTP and CTP levels, decreasing mRNA

315

levels of genes involved in cholesterol and lipid metabolism, and increasing triacylglycerol

316

synthesis and cytosolic lipid droplet formation. Notably, these effects could be reversed by

317

adding either UTP or cytidine. Treatment of cells with pyrazofurin, a known inhibitor of UMP

318

synthase, also increased UTP and CTP levels and promoted lipid droplet formation. Although

319

the mechanisms behind the observed changes in lipid and cholesterol metabolism remain to be

320

elucidated, the cross-talk between pyrimidine biosynthesis and glycerolipid has been clearly

321

demonstrated.

M AN U

SC

RI PT

310

322

(5) Effects of genetic modifications on pyrimidine metabolism.

TE D

323

The pathogenic fungus C. neoformans is responsible for fatal meningitis in

325

immunocompromised patients. A knockout of DHOase (URA4) conferred a growth defect at

326

high temperature, impairing its ability to grow at the physiological temperature of 37°C. RT-

327

PCR revealed that wild type C. neoformans increased expression of genes in both pyrimidine

328

biosynthetic and salvage pathways at restrictive temperatures, in minimal media in the absence

329

of uracil. Under the same conditions, the DHOase knockout mutant exhibited a decrease in

330

expression in biosynthetic genes, while expression of salvage genes remained unchanged. The

331

mutant caused non-lethal infections in mice, nevertheless, the pathogen could be isolated from

332

tissue 40 days post infection (de Gontijo et al., 2014). In a study that sought to explain the

333

synergic effect of 5-fluorocytosine and Amphotericin B combination therapy for cryptococcosis

334

in human immunodeficiency virus (HIV) patients, a C. neoformans knockout of OPRTase

335

(URA5) was constructed. The mutants were found to be hypersensitive to Amphotericin B,

336

probably as a result of a deficiency in the pools of UDP-sugars required for synthesis of the

337

pathogen’s polysaccharide capsule (Banerjee et al., 2014).

AC C

EP

324

12

ACCEPTED MANUSCRIPT

Zameitat and co-workers (2007) deleted the pyr4 gene coding for the mitochondrial

339

DHODase from the basidiomycetous plant pathogen Ustilago maydis. The mutated fungus was

340

auxotrophic for pyrimidines, and exhibited enhanced sensitivity to UV radiation that had been

341

previously reported for mutants deficient in other enzymes of pyrimidine biosynthesis (for

342

example, pyr1). The radiation sensitivity of the pyr4 mutant was completely alleviated by

343

complementation with a human DHODase construct, indicating that the sensitivity is directly

344

related to impairment of de novo synthesis. Importantly, the pyr4 mutant was nonpathogenic in

345

maize, thus de novo synthesis appears to be essential for virulence.

RI PT

338

In the nematode Caenorhabditis elegans, the rad-6 gene, whose mutation results in

347

hypersensitivity to DNA damaging agents, was recently identified as UMP synthase. The

348

hypersensitivity was attributed to the predicted lower concentrations of pyrimidine pools that

349

would decrease their availability for DNA repair. Additionally, mutants exhibited phenotypes

350

associated with defects of proteoglycan synthesis and orotate accumulation, and had reduced

351

lifespans (Merry et al., 2014). When the salvage enzyme uridine kinase (UKase) was knocked

352

down with RNAi in the rad-6 mutant, viability was drastically reduced. In contrast, a UKase

353

knockout in the wild type worm had little effect, suggesting that C. elegans relies more on the de

354

novo pathway than on pyrimidine salvage.

TE D

M AN U

SC

346

A similar situation was found in the apicomplexan parasite T. gondii, where mutants

356

lacking the salvage enzymes UPRTase (Pfefferkorn and Pfefferkorn, 1979) or UPase (Fox and

357

Bzik, 2010) were both replication competent and virulent. In contrast, parasites with knockouts

358

of the T. gondii biosynthetic enzymes CPSase (Fox and Bzik, 2002), or ODCase (Fox and Bzik,

359

2010) were avirulent in mice, and were unable to replicate in cell culture. The growth of both of

360

these knockouts could be rescued in vitro with uracil supplementation. Interestingly, it was not

361

possible to construct a DHODase knockout, even in the presence of uracil supplementation

362

(Hortua Triana et al., 2012). In P. falciparum, transgenic parasites expressing a cytosolic,

363

fumarate-dependent yeast DHODase were found to be resistant to specific inhibitors for the

364

endogenous DHODase, and also exhibited resistance to mitochondrial electron transport chain

365

inhibitors (Painter et al., 2007; Ke et al., 2011), nevertheless, a PfDHODase knockout has not

366

been reported. All of the transgenic strains were able to maintain long-term growth in presence

367

of PfDHODase inhibitors, but some of the strains required decylubiquinone supplementation for

AC C

EP

355

13

ACCEPTED MANUSCRIPT

368

long-term survival in the presence of mitochondrial electron transport chain inhibitors (Ke et al.,

369

2011). In some organisms, genetic manipulations of pyrimidine metabolism lead to the use of

371

alternative pathways. An example is the double knockout of the two alleles of UMP synthase in

372

T. brucei (Ong et al., 2013). While the original mutant parasites were nonvirulent in mice,

373

mutants maintained in long-term tissue culture unexpectedly regained virulence in mice. Uptake

374

studies showed that both original and long-term mutants increased their transport of uracil from

375

the media compared to wild type parasites. Unlike wild type parasites, both original and long-

376

term mutants secreted orotate when grown in pyrimidine-poor environments, however, the

377

latter’s secretion levels were higher. The authors hypothesized that the mutants’ secretion of

378

orotate into the mouse would result in the mouse metabolizing this intermediate into a

379

pyrimidine that the parasite could then transport to supplement its pyrimidine deficiency (Ong et

380

al., 2013).

SC

M AN U

381

RI PT

370

Another example of a microorganism exhibiting adaptability in pyrimidine metabolism is

382

Lachancea kluyveri.

383

pyrimidines:

384

dehydrogenase, which is absent in all fungi), and the newly discovered uracil catabolic pathway

385

(URC) pathway (Andersen et al., 2008). Microarray expression analysis was performed to

386

determine which gene transcripts were affected when different pyrimidine metabolites were used

387

as the sole nitrogen source (Andersson Rasmussen et al., 2014). The URC genes were induced

388

by uracil, while reductive pathway genes (PYD) were induced by dihydrouracil. A new gene

389

coding for the enzyme catalyzing the final step in the URC pathway was identified. In addition,

390

eight new genes encoding putative permeases (transporters) were identified. These permeases

391

appeared to exhibit functional redundancy, since all single and double knockouts strains were

392

found to be viable (Andersson Rasmussen et al., 2014).

This yeast possesses two of the four known catabolic pathways for degradation

(excluding

the

first

enzyme,

dihydropyrimidine

AC C

EP

TE D

reductive

393

Unexpectedly, antisense knockdown of the UMP synthase of S. tuberosum resulted in

394

increased biosynthesis of starch and cell wall in growing tubers (Geigenberger et al., 2005).

395

Transgenic plants were generated by using antisense inhibition of UMP synthase under the

396

control of a tuber-specific promoter. Plant lines with 30 – 85% reduced enzymatic activity in

397

growing tubers were selected. A decrease in UMP synthase activity of ≥ 40% was accompanied

398

by an increase in the expression of CPSase and also by expression increases in the salvage

14

ACCEPTED MANUSCRIPT

enzymes UKase and UPRTase, these in turn, resulted in an increase in the pools of UTP, UDP,

400

and UDP-glucose up to 2-fold higher than wild type. The authors concluded that the wild type

401

uridine nucleotide pools limit the use of sucrose for starch and cell wall synthesis, and they

402

commented that it is surprising that mechanisms for decreasing de novo synthesis have not

403

developed during evolution or in the course of plant breeding, speculating that the pathway must

404

provide advantages under certain conditions (Geigenberger et al., 2005). In a contemporary

405

study on S. tuberosum and tobacco from Zrenner’s lab (Schröder et al., 2005), an antisense

406

strategy was used to produce potato plant lines with ATCase or DHOase proteins levels reduced

407

to < 50% of wild type. These plants showed decreased growth, but exhibited no pleiotropic

408

effects or alterations in the developmental program, and only slight changes in pyrimidine

409

nucleotide levels in mature tissues. Carbohydrate levels were decreased in roots, and increased

410

in leaves. Plants with < 10% of wild type DHOase protein eventually died without flowering or

411

tuberization. Tobacco plant lines with decreased UMP synthase expression were produced but

412

exhibited no phenotypic differences from wild type, however, the authors commented that the

413

reduction in the protein levels may not have been sufficient to cause such differences. They

414

concluded that the response of the plant to a moderate decrease in pyrimidine biosynthesis was to

415

decrease its growth rate in accordance with the size of the available nucleotide pool, however,

416

they noted that the size of the pool may depend on the type of tissue and its developmental stage

417

(Schröder et al., 2005). Support for this idea was provided by a later study, which demonstrated

418

dynamic regulation of the PYRB-encoded ATCase protein during Arabidopsis thaliana growth

419

and development as shown by monitoring the activity of a reporter gene under the control of the

420

PYRB promoter (Chen and Slocum, 2008). Similarly, Arabidopsis lines with RNAi-mediated 70

421

– 95% knockdowns of PYRB transcripts and concomitant reduced ATCase protein levels had

422

delayed growth and development compared with wild type plants (Chen and Slocum, 2008).

423

Plants having only 5% of the wild type expression produced seeds with extremely poor viability.

424

This Arabidopsis study is in general agreement with the results of Geigenberger and co-workers

425

(Geigenberger et al., 2005), for example, knockdown of a pyrimidine biosynthetic enzyme, in

426

this case ATCase, resulted in increased expression of another biosynthetic enzyme, in this case

427

UMP synthase, and increased expression of the salvage enzyme UPRTase (Chen and Slocum,

428

2008). The conclusions that can be drawn from these three studies are that plants dynamically

429

regulate synthetic and salvage pathways according to pyrimidine availability, and a decrease of

AC C

EP

TE D

M AN U

SC

RI PT

399

15

ACCEPTED MANUSCRIPT

430

flux through the former simply leads to a slower growth rate but does not result in phenotypic

431

abnormalities. However, the synthetic pathway is clearly essential, because decreasing activity

432

below certain thresholds results in loss of viability. Genetic manipulations of pyrimidine metabolism in multicellular organisms often have

434

complex effects that are not easy to dissect. Recent studies on UPase in mice illustrate this point

435

(Le et al., 2013; Urasaki et al., 2014). A knock-in of UPase resulted in enzyme overexpression

436

and a 25-fold increase in activity. The livers of the mice exhibited a 13-fold decrease in uridine

437

levels, and more than 2-fold higher levels of beta-alanine. The mice were found to suffer from

438

hepatic vesicular steatosis, which could be suppressed by administration of uridine or by

439

inhibition of dihydropyrimidine dehydrogenase. Changes in liver protein acetylation upon

440

administration of uridine were observed in mice and in primary hepatocyte cultures. Testing of

441

inhibitors and metabolites of de novo pyrimidine biosynthesis showed that brequinar, an inhibitor

442

of DHODase, and its product orotate, also caused hepatic vesicular steatosis (Le et al., 2013).

443

The authors attributed the effect of orotate to inhibition of DHODase and suggested that this

444

compromised mitochondrial respiration in the liver. Nevertheless, orotate is known to affect

445

expression levels of enzymes involved in lipid metabolism (Jung et al., 2011).

446

important to point out that the contribution of DHODase activity to electron transport flux or to

447

the function of respiratory complexes is still unknown. Furthermore, human genetic deficiencies

448

in UMP synthase, which lead to high levels of orotate causing orotic aciduria, and deficiencies in

449

DHODase (Miller syndrome), have very different clinical features (Balasubramaniam et al.,

450

2014). In another study, mice with a knockout of UPase (UPase-/-) were found to have uridine

451

increases of 6-fold in plasma and 3-fold in the liver (Le et al., 2013; Urasaki et al., 2014). They

452

exhibited a decrease in insulin-stimulated blood glucose removal, and an activation of the heme-

453

deficiency response pathway. Interestingly, although these last two effects were also observed in

454

wild type mice treated with short term supplementation of uridine, additional effects, for

455

example increased liver protein glycosylation and reduced phosphorylation of insulin signaling

456

proteins, were not observed in the knockout mice.

It is also

AC C

EP

TE D

M AN U

SC

RI PT

433

457 458

(6) Concluding remarks.

459

In this review, we have presented several studies on pyrimidine metabolism in pathogen-

460

infected hosts. During their evolutionary development, pathogens have developed strategies for

16

ACCEPTED MANUSCRIPT

manipulating pyrimidine metabolism in the organisms they infect, thereby generating cellular

462

host environments that provide optimum levels of pyrimidine metabolites necessary for pathogen

463

proliferation. We would like to learn how these consummate manipulators of host cell

464

metabolism alter pyrimidine levels for their own ends. Perhaps such knowledge could be applied

465

to the ongoing challenge of understanding how pyrimidine inhibitors or antimetabolites can be

466

used to achieve a desired effect, for example to kill viruses, parasites, or other pathogens. Several

467

obstacles have impeded researchers from achieving these goals. A problem inherent in the study

468

of pathogen-infected hosts is that it is difficult or impossible in many cases to determine which

469

organism the metabolites originated from. Another problem is that information on the cellular

470

system under study is often incomplete, for example, expression data may be readily available,

471

but information on mRNA stability, metabolite levels, or enzyme activities may be lacking. The

472

integration and analysis of available data from new strategies, such as genomics, transcriptomics,

473

proteomics, metabolomics, is beginning to permit the construction of a more complete picture. It

474

should be noted that although these new techniques will permit a more comprehensive

475

description of the effect of an inhibitor or a genetic manipulation on the cell, the mechanisms

476

which led to that effect may still not be understood. Despite extensive biochemical knowledge

477

about pyrimidine metabolism, a global vision of its dynamics is only starting to emerge.

478 479

Acknowledgements

TE D

M AN U

SC

RI PT

461

This work was supported in part by the Universidad de los Andes Faculty of Sciences, by

481

basic sciences project P13.700022.007 from the University of los Andes Office of Research, by

482

COLCIENCIAS grant # 120-4521-28532. We would like to give a special acknowledgement to

483

Colciencias (Programa de Apoyo a Doctorados Nacionales) for the financial support of the Ph.D.

484

studies of M.F.G. and H.Y.N.-O. We regret that some important findings were not included here

485

due to the limited space.

487

AC C

486

EP

480

488

References

489 490 491

Ali, J.A.M., Creek, D.J., Burgess, K., Allison, H.C., Field, M.C., Mäser, P., De Koning, H.P., 2013. Pyrimidine salvage in Trypanosoma brucei bloodstream forms and the trypanocidal action of halogenated pyrimidines. Mol. Pharmacol. 83, 439–453.

17

ACCEPTED MANUSCRIPT

Andersen, G., Björnberg, O., Polakova, S., Pynyaha, Y., Rasmussen, A., Møller, K., Hofer, A., Moritz, T., Sandrini, M.P.B., Merico, A.-M., Compagno, C., Akerlund, H.-E., Gojković, Z., Piskur, J., 2008. A second pathway to degrade pyrimidine nucleic acid precursors in eukaryotes. J. Mol. Biol. 380, 656–666.

496 497 498 499

Andersson Rasmussen, A., Kandasamy, D., Beck, H., Crosby, S.D., Björnberg, O., Schnackerz, K.D., Piškur, J., 2014. Global expression analysis of the yeast Lachancea (Saccharomyces) kluyveri reveals new URC genes involved in pyrimidine catabolism. Eukaryot. Cell 13, 31– 42.

500 501

Angeletti, P.C., Engler, J.A., 1998. Adenovirus preterminal protein binds to the CAD enzyme at active sites of viral DNA replication on the nuclear matrix. J. Virol. 72, 2896–2904.

502 503

Balasubramaniam, S., Duley, J., Christodoulou, J., 2014. Inborn errors of pyrimidine metabolism: clinical update and therapy. J. Inherit. Metab. Dis. 37, 687–698.

504 505

Banerjee, D., Burkard, L., Panepinto, J.C., 2014. Inhibition of nucleotide biosynthesis potentiates the antifungal activity of amphotericin B. PLoS One 9, e87246.

506 507 508 509

Bardeleben, C., Sharma, S., Reeve, J.R., Bassilian, S., Frost, P., Hoang, B., Shi, Y., Lichtenstein, A., 2013. Metabolomics identifies pyrimidine starvation as the mechanism of 5aminoimidazole-4-carboxamide-1-β-riboside-induced apoptosis in multiple myeloma cells. Mol. Cancer Ther. 12, 1310–1321.

510 511 512

Barry, R.M., Bitbol, A., Lorestani, A., Charles, E.J., Habrian, C.H., Hansen, J.M., Li, H., Baldwin, E.P., Wingreen, N.S., Kollman, J.M., Gitai, Z., 2014. Large-scale filament formation inhibits the activity of CTP synthetase. Elife 3, e03638.

513 514

Ben-Sahra, I., Howell, J.J., Asara, J.M., Manning, B.D., 2013. Stimulation of de novo pyrimidine synthesis by growth signaling through mTOR and S6K1. Science 339, 1323–1328.

515 516 517

Carcamo, W.C., Calise, S.J., von Mühlen, C. a, Satoh, M., Chan, E.K.L., 2014. Molecular cell biology and immunobiology of mammalian rod/ring structures. Int. Rev. Cell Mol. Biol. 308, 35–74.

518 519

Chauhan, M., Kumar, R., 2013. Medicinal attributes of pyrazolo[3,4-d]pyrimidines: a review. Bioorg. Med. Chem. 21, 5657–5668.

520 521 522

Chen, C.T., Slocum, R.D., 2008. Expression and functional analysis of aspartate transcarbamoylase and role of de novo pyrimidine synthesis in regulation of growth and development in Arabidopsis. Plant Physiol. Biochem. 46, 150–159.

523 524 525

Chen, K., Zhang, J., Tastan, Ö.Y., Deussen, Z.A., Siswick, M.Y.-Y., Liu, J.-L., 2011. Glutamine analogs promote cytoophidium assembly in human and Drosophila cells. J. Genet. Genomics 38, 391–402.

AC C

EP

TE D

M AN U

SC

RI PT

492 493 494 495

18

ACCEPTED MANUSCRIPT

Christopherson, R.I., Jones, M.E., 1979. Interconversion of carbamayl-L-aspartate and Ldihydroorotate by dihydroorotase from mouse Ehrlich ascites carcinoma. J. Biol. Chem. 254, 12506–12512.

529 530

Cox, R.A., 1968. Macromolecular structure and properties of ribonucleic acids. Q. Rev. Chem. Soc. 22, 499–526.

531 532 533

De Gontijo, F.A., Pascon, R.C., Fernandes, L., Machado, J., Alspaugh, J.A., Vallim, M.A., 2014. The role of the de novo pyrimidine biosynthetic pathway in Cryptococcus neoformans high temperature growth and virulence. Fungal Genet. Biol. 70, 12–23.

534 535

Evans, D.R., Guy, H.I., 2004. Mammalian pyrimidine biosynthesis: fresh insights into an ancient pathway. J. Biol. Chem. 279, 33035–33038.

536 537

Fox, B.A., Bzik, D.J., 2002. De novo pyrimidine biosynthesis is required for virulence of Toxoplasma gondii. Nature 415, 926–929.

538 539 540

Fox, B.A., Bzik, D.J., 2010. Avirulent uracil auxotrophs based on disruption of orotidine-5′monophosphate decarboxylase elicit protective immunity to Toxoplasma gondii. Infect. Immun. 78, 3744–3752.

541 542 543

García-Bayona, L., Garavito, M.F., Lozano, G.L., Vasquez, J.J., Myers, K., Fry, W.E., Bernal, A., Zimmermann, B.H., Restrepo, S., 2014. De novo pyrimidine biosynthesis in the oomycete plant pathogen Phytophthora infestans. Gene 537, 312–321.

544 545 546 547 548

Geigenberger, P., Regierer, B., Nunes-Nesi, A., Leisse, A., Urbanczyk-Wochniak, E., Springer, F., van Dongen, J.T., Kossmann, J., Fernie, A.R., 2005. Inhibition of de novo pyrimidine synthesis in growing potato tubers leads to a compensatory stimulation of the pyrimidine salvage pathway and a subsequent increase in biosynthetic performance. Plant Cell 17, 2077–2088.

549 550 551 552

Hortua Triana, M.A., Huynh, M.-H., Garavito, M.F., Fox, B.A., Bzik, D.J., Carruthers, V.B., Löffler, M., Zimmermann, B.H., 2012. Biochemical and molecular characterization of the pyrimidine biosynthetic enzyme dihydroorotate dehydrogenase from Toxoplasma gondii. Mol. Biochem. Parasitol. 184, 71–81.

553 554 555

Hu, J., Locasale, J.W., Bielas, J.H., O’Sullivan, J., Sheahan, K., Cantley, L.C., Vander Heiden, M.G., Vitkup, D., 2013. Heterogeneity of tumor-induced gene expression changes in the human metabolic network. Nat. Biotechnol. 31, 522–529.

556 557

Huang, M., Graves, L.M., 2003. De novo synthesis of pyrimidine nucleotides; emerging interfaces with signal transduction pathways. Cell. Mol. Life Sci. 60, 321–336.

558 559

Hyde, J.E., 2007. Targeting purine and pyrimidine metabolism in human apicomplexan parasites. Curr. Drug Targets 8, 31–47.

AC C

EP

TE D

M AN U

SC

RI PT

526 527 528

19

ACCEPTED MANUSCRIPT

Jain, K.S., Chitre, T.S., Miniyar, P.B., Kathiravan, M.K., Bendre, V.S., Veer, V.S., Shahane, S.R., Shishoo, C.J., 2006. Biological and medicinal significance of pyrimidines. Curr. Sci. 90, 793–803.

563 564

Jones, M.E., 1980. Pyrimidine nucleotide biosynthesis in animals: genes, enzymes, and regulation of UMP biosynthesis. Annu. Rev. Biochem. 49, 253–279.

565 566

Jung, E.-J., Kwon, S.-W., Jung, B.-H., Oh, S.-H., Lee, B.-H., 2011. Role of the AMPK/SREBP-1 pathway in the development of orotic acid-induced fatty liver. J. Lipid Res. 52, 1617–1625.

567 568 569

Jung, E.-J., Lee, K.-Y., Lee, B.-H., 2012. Proliferating effect of orotic acid through mTORC1 activation mediated by negative regulation of AMPK in SK-Hep1 hepatocellular carcinoma cells. J. Toxicol. Sci. 37, 813–821.

570 571 572

Katahira, R., Ashihara, H., 2006. Dual function of pyrimidine metabolism in potato (Solanum tuberosum) plants: pyrimidine salvage and supply of beta-alanine to pantothenic acid synthesis. Physiol. Plant. 127, 38–43.

573 574 575

Ke, H., Morrisey, J.M., Ganesan, S.M., Painter, H.J., Mather, M.W., Vaidya, A.B., 2011. Variation among Plasmodium falciparum strains in their reliance on mitochondrial electron transport chain function. Eukaryot. Cell 10, 1053–1061.

576 577 578

Kusch, H., Engelmann, S., Bode, R., Albrecht, D., Morschhäuser, J., Hecker, M., 2008. A proteomic view of Candida albicans yeast cell metabolism in exponential and stationary growth phases. Int. J. Med. Microbiol. 298, 291–318.

579 580 581

Le, T.T., Ziemba, A., Urasaki, Y., Hayes, E., Brotman, S., Pizzorno, G., 2013. Disruption of uridine homeostasis links liver pyrimidine metabolism to lipid accumulation. J. Lipid Res. 54, 1044–1057.

582 583

Liu, J.-L., 2011. The enigmatic cytoophidium: compartmentation of CTP synthase via filament formation. Bioessays 33, 159–164.

584 585 586

Loh, K.D., Gyaneshwar, P., Markenscoff Papadimitriou, E., Fong, R., Kim, K.-S., Parales, R., Zhou, Z., Inwood, W., Kustu, S., 2006. A previously undescribed pathway for pyrimidine catabolism. Proc. Natl. Acad. Sci. U. S. A. 103, 5114–5119.

587 588

Merry, A., Qiao, M., Hasler, M., Kuwabara, P.E., 2014. RAD-6: pyrimidine synthesis and radiation sensitivity in Caenorhabditis elegans. Biochem. J. 458, 343–353.

589 590

Munger, J., Bajad, S.U., Coller, H.A., Shenk, T., Rabinowitz, J.D., 2006. Dynamics of the cellular metabolome during human cytomegalovirus infection. PLoS Pathog. 2, e132.

591 592

Nara, T., Hshimoto, T., Aoki, T., 2000. Evolutionary implications of the mosaic pyrimidinebiosynthetic pathway in eukaryotes. Gene 257, 209–222.

AC C

EP

TE D

M AN U

SC

RI PT

560 561 562

20

ACCEPTED MANUSCRIPT

Olszewski, K.L., Morrisey, J.M., Wilinski, D., Burns, J.M., Vaidya, A.B., Rabinowitz, J.D., Llinás, M., 2009. Host-parasite interactions revealed by Plasmodium falciparum metabolomics. Cell Host Microbe 5, 191–199.

596 597 598

Ong, H.B., Sienkiewicz, N., Wyllie, S., Patterson, S., Fairlamb, A.H., 2013. Trypanosoma brucei (UMP synthase null mutants) are avirulent in mice, but recover virulence upon prolonged culture in vitro while retaining pyrimidine auxotrophy. Mol. Microbiol. 90, 443–455.

599 600

Painter, H.J., Morrisey, J.M., Mather, M.W., Vaidya, A.B., 2007. Specific role of mitochondrial electron transport in blood-stage Plasmodium falciparum. Nature 446, 88–91.

601 602

Patel, B.N., West, T.P., 1987. Oxidative catabolism of uracil by Enterobacter aerogenes. FEMS Microbiol. Lett. 40, 33–36.

603 604

Pfefferkorn, E.R., Pfefferkorn, L.C., 1979. Quantitative studies of the mutagenesis of Toxoplasma gondii. J. Parasitol. 65, 364–370.

605 606 607

Poirier, S., Samami, S., Mamarbachi, M., Demers, A., Chang, T.Y., Vance, D.E., Hatch, G.M., Mayer, G., 2014. The epigenetic drug 5-azacytidine interferes with cholesterol and lipid metabolism. J. Biol. Chem. 289, 18736–18751.

608 609 610

Robitaille, A.M., Christen, S., Shimobayashi, M., Cornu, M., Fava, L.L., Moes, S., PrescianottoBaschong, C., Sauer, U., Jenoe, P., Hall, M.N., 2013. Quantitative phosphoproteomics reveal mTORC1 activates de novo pyrimidine synthesis. Science 339, 1320–1323.

611 612 613

Sahoo, S., Aurich, M.K., Jonsson, J.J., Thiele, I., 2014. Membrane transporters in a human genome-scale metabolic knowledgebase and their implications for disease. Front. Physiol. 5, 91.

614 615

Schröder, M., Giermann, N., Zrenner, R., 2005. Functional analysis of the pyrimidine de novo synthesis pathway in solanaceous species. Plant Physiol. 138, 1926–1938.

616 617

Sharma, V., Chitranshi, N., Agarwal, A.K., 2014. Significance and biological importance of pyrimidine in the microbial world. Int. J. Med. Chem. 2014, 202784.

618 619 620

Shock, J.L., Fischer, K.F., DeRisi, J.L., 2007. Whole-genome analysis of mRNA decay in Plasmodium falciparum reveals a global lengthening of mRNA half-life during the intraerythrocytic development cycle. Genome Biol. 8, R134.

621 622

Sigoillot, F.D., Berkowski, J.A., Sigoillot, S.M., Kotsis, D.H., Guy, H.I., 2003. Cell cycledependent regulation of pyrimidine biosynthesis. J. Biol. Chem. 278, 3403–3409.

623 624 625

Sreedharan, S., Shaik, J.H.A., Olszewski, P.K., Levine, A.S., Schiöth, H.B., Fredriksson, R., 2010. Glutamate, aspartate and nucleotide transporters in the SLC17 family form four main phylogenetic clusters: evolution and tissue expression. BMC Genomics 11, 17.

AC C

EP

TE D

M AN U

SC

RI PT

593 594 595

21

ACCEPTED MANUSCRIPT

Stasolla, C., Loukanina, N., Yeung, E.C., Thorpe, T. a, 2004. Alterations in pyrimidine nucleotide metabolism as an early signal during the execution of programmed cell death in tobacco BY-2 cells. J. Exp. Bot. 55, 2513–2522.

629 630

Tiedje, K.E., Stevens, K., Barnes, S., Weaver, D.F., 2010. Beta-alanine as a small molecule neurotransmitter. Neurochem. Int. 57, 177–188.

631 632

Urasaki, Y., Pizzorno, G., Le, T.T., 2014. Uridine affects liver protein glycosylation, insulin signaling, and heme biosynthesis. PLoS One 9, e99728.

633 634 635

Vastag, L., Koyuncu, E., Grady, S.L., Shenk, T.E., Rabinowitz, J.D., 2011. Divergent effects of human cytomegalovirus and herpes simplex virus-1 on cellular metabolism. PLoS Pathog. 7, e1002124.

636 637

Vogels, G and van der Drift, C., 1976. Degradation of purines and pyrimidines by microorganisms. Bacteriol. Rev. 40, 403–468.

638 639

Webb, M.E., Smith, A.G., Abell, C., 2004. Biosynthesis of pantothenate. Nat. Prod. Rep. 21, 695–721.

M AN U

SC

RI PT

626 627 628

640

Fig. 1. Overview of human pyrimidine metabolic pathways.

642

Transporters important for pyrimidine metabolism are shown at the top of the figure. Human

643

cells can use transporters from the solute carrier family (SCL17) to exchange amino acids

644

(aspartate and glutamate) and nucleotides (Sreedharan et al., 2010), while nucleobases can exit or

645

enter the cell via the equilibrative nucleoside transporters (ENTs) and the concentrative

646

nucleoside transporters (CNT1) (Sahoo et al., 2014). Transporters of the ABC transport family

647

import nucleotides and nucleosides into the cell. The de novo pyrimidine biosynthetic pathway

648

includes six enzymatic steps (far left, purple). The first three steps are catalyzed by a

649

multifunctional protein called CAD and the last two steps by a bifunctional enzyme, UMP

650

synthase.

651

transferases (cyan). Nucleosides may be phosphorylated to form nucleoside monophosphates,

652

diphosphates and triphosphates by nucleoside kinases (pink and orange). Nucleoside

653

diphosphates can be transformed into deoxynucleosides by the ribonucleotide reductase system.

AC C

EP

TE D

641

Inside the cells, nucleobases are converted to nucleosides by phosphoribosyl

22

ACCEPTED MANUSCRIPT

The lower part of the figure shows that pyrimidine nucleotides are the main precursors for

655

synthesizing lipids (blue) and UDP-sugars (gray), and are also essential for the formation of

656

RNA (green) and DNA (brown). Pyrimidines are degraded to ß-alanine or ß-aminoisobutyrate

657

(far right, yellow). The latter can be further degraded to methylmalonate-semialdehyde which is

658

an intermediate in valine degradation. Enzymes are denoted by their Enzyme Commission (EC)

659

numbers: Dihydropyrimidine dehydrogenase (1.3.1.2); dihydroorotate dehydrogenase (1.3.5.2);

660

ribonucleoside-diphosphate reductase (1.17.4.1); thioredoxin-disulfide reductase (1.8.1.9);

661

thymidylate synthase (2.1.1.45); aspartate transcarbamoylase (2.1.3.2); uridine phosphorylase

662

(2.4.2.3); thymidine phosphorylase (2.4.2.4); uracil phosphoribosyltransferase (2.4.2.9); orotate

663

phosphoribosyltransferase (2.4.2.10); thymidine kinase (2.7.1.21); uridine kinase (2.7.1.48);

664

deoxycytidine kinase (2.7.1.74); nucleoside diphosphate kinase (2.7.4.6); thymidylate kinase

665

(2.7.4.9); adenylate kinase G (2.7.4.10); UMP/CMP kinase (2.7.4.14); RNA polymerase

666

(2.7.7.6); DNA polymerase (2.7.7.7); UDP-glucose pyrophosphorylase (2.7.7.9); ethanolamine-

667

phosphate cytidylyltransferase (2.7.7.14); choline-phosphate cytidylyltransferase (2.7.7.15);

668

UTP:N-acetyl-a-D-glucosamine-1-phosphate

669

cytidylyltransferase

670

pyrophosphatase (3.6.1.19); beta-ureidopropionase (3.5.1.6); dihydropyrimidinase (3.5.2.2);

671

dihydroorotase

672

diphosphatase (3.6.1.6); dCTP pyrophosphatase 1 (3.6.1.12); deoxyuridine triphosphatase

673

(3.6.1.23); orotidine 5-phosphate decarboxylase (4.1.1.23); cytidine triphosphate synthase

674

(6.3.4.2); carbamoyl phosphate synthase (6.3.5.5). Abbreviations: ADP, adenosine diphosphate;

675

Asp, aspartate; ATP, adenosine triphosphate; ß-Ala, ß-alanine; ß-AIB, ß-aminoisobutyrate; ß-

676

UIB, ß-ureidoisobutyrate; ß-UP, ß-ureidopropionate; CDP, cytidine-5-diphosphate; CDP-Cho,

677

CDP-choline; CDP-Etn, CDP-ethanolamine; CMP, cytidine-5-monophosphate; CTP, cytidine-5-

TE D

M AN U

SC

RI PT

654

EP

(2.7.7.41);

AC C

(3.5.2.3);

cytidine

uridylyltransferase

5′-nucleotidase

deaminase

(3.1.3.5);

(3.5.4.5);

(2.7.7.23);

phosphatidate

nucleoside

triphosphate

apyrase

(3.6.1.5);

nucleoside

23

ACCEPTED MANUSCRIPT

678

triphosphate; dCDP, deoxy-cytidine-5-diphosphate; dCMP, deoxy-cytidine-5-monophosphate;

679

dCTP, deoxy-cytidine-5-triphosphate; DHF, dihydrofolate; DHT, dihydrothymine; DHU,

680

dihydrouracil;

681

monophosphate; dTTP, deoxy-thymidine-5-triphosphate; dUDP, deoxy-uridine-5-diphosphate;

682

dUMP, deoxy-uridine-5-monophosphate; dUTP, deoxy-uridine-5-triphosphate; Glu, glutamate;

683

Glu-1-P, glucose-1-phosphate; Gly, glycine; NADPH, nicotinamide adenine dinucleotide

684

phosphate; OMP, orotidine-5-monophosphate; Pi, inorganic phosphate; PPi, pyrophosphate;

685

PRPP, phosphoribosyl pyrophosphate; Ser, serine; UDP, uridine-5-diphosphate; UMP, uridine-5-

686

monophosphate; UQ, ubiquinone; UTP, uridine-5-triphosphate; THF, tetrahydrofolate.

687

diagram is based on data from the Kyoto Encyclopedia of Genes and Genomes (KEGG)

688

Pathway, hsa00240 (http://www.genome.jp/kegg-bin/show_pathway?map00240). ChemAxon

689

software was employed to build the chemical structures.

690 691

Fig. 2. A heat map illustrating changes in the expression of the genes of pyrimidine metabolism

692

in pathogenic eukaryotes.

693

A graphical heat map representation was constructed to illustrate changes (log2) in expression

694

among two contrasting time points or conditions important in the infection process of selected

695

plant and animal pathogens. Only RNA Sequencing (RNA-seq) data were used for its

696

construction since these represent the quantity of the transcripts at a given moment or condition.

697

Red and green color intensities indicate increases and decreases, respectively, in the gene

698

expression profile. Gray indicates the absence of the gene. Toxoplasma gondii: data were

699

obtained for tachyzoites of the RH strain from infected human foreskin fibroblasts from ToxoDB

700

13.0. Time points were early in infection, 2 hpi, and late in infection, 20 hpi. Plasmodium

701

falciparum: data were obtained for strain 3D7 during intra-erythrocytic development from

702

PlasmoDB 9.0. Selected time points were 20 hpi (trophozoite) and 40 hpi (schizont).

703

Cryptosporidium parvum: data were obtained for sporozoites of isolate IPZ:CH-Crypto_K6769

704

infection of CT-8 cells from CryptoDB 7.0. The time points selected were infected cells

705

incubated for 48 hpi and 96 hpi. Trypanosoma brucei: data were obtained for strain Lister 427

deoxy-thymidine-5-diphosphate;

dTMP,

deoxy-thymidine-5-

The

AC C

EP

TE D

M AN U

SC

RI PT

dTDP,

24

ACCEPTED MANUSCRIPT

from TriTrypDB 8.1. The comparison was generated between proliferative bloodstream stage in

707

the mammalian host, and the procyclic form, from which it differentiates, in the midgut of the

708

tsetse fly. Cryptococcus neoformans: data were obtained for variety grubii H99 from wild-type

709

data from the FungiDB 3.1. The comparison was generated between capsule inducing (37°C with

710

CO2) and non-inducing (30°C) conditions. Colletotricum higginsianum: a pathogenic plant

711

fungus, data were obtained for the in planta stages during infection of Arabidopsis thaliana leaf

712

sheaths from the Gene Expression Omnibus GSE33683. Selected time points were 22 hpi,

713

representing the earliest stage of biotrophy with penetrating apressoria, and 60 hpi in the

714

necrotrophic phase. Phytophthora ramorum: data were obtained from the FungiDB 3.1. The

715

comparison was generated between (sporangia+zoospores) germination stages and non-

716

germinating mycelia grown in tomato broth. In addition, we present expression data for an

717

infected host, Mus musculus (mouse macrophages): data were obtained for primary bone

718

marrow-derived macrophages from C57BL/6 mice infected T. gondii RH strain from HostDB

719

1.1. Time points were uninfected macrophages and infected macrophages 22 hpi. Viewing was

720

done with Mev 4.8.1.

M AN U

SC

RI PT

706

AC C

EP

TE D

721

25

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT

Supplementary Table 1. Analyses of the genes of pyrimidine metabolism in selected eukaryotic genomes.

✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓

✓ ✓ ✓ ✓ ✓ ✓ ✓

✓ ✓ ✓ ✓ ✓ ✓ ✓

✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓

Low mRNA expression levels Exclusively for IMP/GMP Replaced by 3.6.1.6 Nucleoside diphosphate phosphatase (no information about whether it is purine or pyrimidine specific) Gene exists but is inactive Absent

✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓

✓ ✓ ✓ ✓ ✓ ✓

✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓

2.4.2.3 Uridine phosphorylase

3.5.4.5 Cytidine deaminase

3.1.3.5 5'-nucleotidase

3.6.1.5 Apyrase

3.6.1.12 dCTP pyrophosphatase

2.7.1.21 Thymidine kinase

2.7.4.9 Thymidylate kinase

2.7.1.74 Deoxycytidine kinase

2.7.1.48 Uridine kinase

2.7.4.6 Nucleoside diphosphate kinase

2.7.4.14 UMP/CMP kinase

✓ ✓ ✓ ✓ ✓

3.6.1.19 Nucleoside triphosphate pyrophosphatase

✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓

RI PT

✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓

SC

2.1.1.45 Thymidylate synthase ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓

2.4.2.10 Orotate phosphoribosyltransferase

1.3.5.2 Dihydroorotate dehydrogenase

6.3.4.2 Cytidine triphosphate synthase

✓ ✓ ✓

✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓

✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓ ✓

TE D

3.5.2.3 Dihydroorotase ✓ ✓ ✓ ✓ ✓ ✓ ✓

EP

✓ ✓ ✓ ✓

4.1.1.23 Orotidine 5-phosphate decarboxylase

✓ ✓ ✓

✓ ✓ ✓ ✓

2.1.3.2 Aspartate transcarbamoylase

6.3.5.5 Carbamoyl phosphate synthase

3.5.1.6 Beta-ureidopropionase

✓ ✓ ✓ ✓

AC C

Cryptosporidium parvum Plasmodium falciparum Toxoplasma gondii Trypanosoma brucei Phytophthora ramorum Candida albicans Ustilago maydis Cryptococcus neoformans Caenorhabditis elegans Arabidopsis thaliana Homo sapiens

3.5.2.2 Dihydropyrimidinase

Organism

1.3.1.2 Dihydropyrimidine dehydrogenase

E.C number - Enzyme

2.4.2.9 Uracil phosphoribosyltransferase

Salvage & Modifications 1.17.4.1 Ribonucleoside-diphosphate reductase

Synthesis

M AN U

Catabolism

✓ ✓ ✓ ✓ ✓

ACCEPTED MANUSCRIPT

Supplementary Table 1. Analyses of the genes of pyrimidine metabolism in selected eukaryotic genomes.

AC C

EP

TE D

M AN U

SC

RI PT

The table highlights the differences in the genomic metabolic repertoire of the pyrimidine pathways among the eukaryotes mentioned in this review. Presence or absence of the sequences of the enzymes of pyrimidine metabolism was evaluated for each organism using KEGG, accessing by gene code, name, and EC number. A second validation was performed with the same criteria but using instead the data available in each organism’s specific genome database. For genes not present, possible misannotation problems were further evaluated by analyzing the genomes to look for hypothetical proteins that shared the conserved protein domain of the target protein. A check mark represents the presence of coding sequence for a protein. Enzymes are grouped as pathways of catabolism, synthesis, salvage and modification. Enzyme numbers are the same as in Figure 1. Specific genomic data were retrieved from: Cryptosporidium parvum: http://cryptodb.org; Plasmodium falciparum: http://plasmodb.org; Toxoplasma gondii: http://toxodb.org; Trypanosoma brucei: http://tritrypdb.org; Phytophthora ramorum: http://genome.jgi-psf.org/; Candida albicans: http://www.candidagenome.org; Ustilago maydis: http://www.broadinstitute.org; Cryptococcus neoformans: http://www.broadinstitute.org; Caenorhabditis elegans: http://www.wormbase.org; Arabidopsis thaliana: http://www.arabidopsis.org/index.jsp; Homo sapiens: http://www.genome.jp/kegg-bin/show_pathway?org_name=hsa&mapno=00240&mapscale=&show_description=show.

Pyrimidine Metabolism: Dynamic and Versatile Pathways in Pathogens and Cellular Development.

The importance of pyrimidines lies in the fact that they are structural components of a broad spectrum of key molecules that participate in diverse ce...
3MB Sizes 1 Downloads 9 Views