conversion of prps hexamer to monomer by ampk-mediated … · 94 | cancer discovery january 2018...

15
94 | CANCER DISCOVERY JANUARY 2018 www.aacrjournals.org Conversion of PRPS Hexamer to Monomer by AMPK-Mediated Phosphorylation Inhibits Nucleotide Synthesis in Response to Energy Stress Xu Qian 1 , Xinjian Li 1 , Lin Tan 2 , Jong-Ho Lee 1 , Yan Xia 1 , Qingsong Cai 1 , Yanhua Zheng 1 , Hongxia Wang 3 , Philip L. Lorenzi 2 , and Zhimin Lu 1,4,5 RESEARCH ARTICLE ABSTRACT Tumors override energy stress to grow. However, how nucleotide synthesis is regulated under energy stress is unclear. We demonstrate here that glucose dep- rivation or hypoxia results in the AMPK-mediated phosphorylation of phosphoribosyl pyrophosphate synthetase 1 (PRPS1) S180 and PRPS2 S183, leading to conversion of PRPS hexamers to monomers and thereby inhibiting PRPS1/2 activity, nucleotide synthesis, and nicotinamide adenine dinucleotide (NAD) production. Knock-in of nonphosphorylatable PRPS1/2 mutants, which have uninhibited activ- ity, in brain tumor cells under energy stress exhausts cellular ATP and NADPH and increases reactive oxygen species levels, thereby promoting cell apoptosis. The expression of those mutants inhibits brain tumor formation and enhances the inhibitory effect of the glycolysis inhibitor 2-deoxy-D-glucose on tumor growth. Our findings highlight the significance of recalibrating tumor cell metabolism by fine- tuning nucleotide and NAD synthesis in tumor growth. SIGNIFICANCE: Our findings elucidate an instrumental function of AMPK in direct regulation of nucleic acid and NAD synthesis in tumor cells in response to energy stress. AMPK phosphorylates PRPS1/2, converts PRPS1/2 hexamers to monomers, and inhibits PRPS1/2 activity and subsequent nucleotide and NAD synthesis to maintain tumor cell growth and survival. Cancer Discov; 8(1); 94–107. ©2017 AACR. 1 Brain Tumor Center and Department of Neuro-Oncology, The University of Texas MD Anderson Cancer Center, Houston, Texas. 2 Department of Bioin- formatics and Computational Biology, and The Proteomics and Metabo- lomics Core Facility, The University of Texas MD Anderson Cancer Center, Houston, Texas. 3 The Institute of Cell Metabolism and Disease, Shanghai Key Laboratory of Pancreatic Cancer, Shanghai General Hospital, School of Medicine, Shanghai Jiaotong University, Shanghai, China. 4 Department of Molecular and Cellular Oncology, The University of Texas MD Anderson Cancer Center, Houston, Texas. 5 Cancer Biology Program, MD Anderson Cancer Center UT Health Graduate School of Biomedical Sciences, The University of Texas, Houston, Texas. Note: Supplementary data for this article are available at Cancer Discovery Online (http://cancerdiscovery.aacrjournals.org/). X. Qian and X. Li contributed equally to this article. Corresponding Author: Zhimin Lu, The University of Texas MD Anderson Cancer Center, 1515 Holcombe Boulevard, Unit 1002, Houston, TX 77030. Phone: 713-834-6231; Fax: 713-834-6230; E-mail: zhiminlu@mdanderson. org doi: 10.1158/2159-8290.CD-17-0712 ©2017 American Association for Cancer Research. Research. on October 13, 2020. © 2018 American Association for Cancer cancerdiscovery.aacrjournals.org Downloaded from Published OnlineFirst October 26, 2017; DOI: 10.1158/2159-8290.CD-17-0712

Upload: others

Post on 01-Aug-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Conversion of PRPS Hexamer to Monomer by AMPK-Mediated … · 94 | CANCER DISCOVERY January 2018 Conversion of PRPS Hexamer to Monomer by AMPK-Mediated Phosphorylation Inhibits Nucleotide

94 | CANCER DISCOVERY January 2018 www.aacrjournals.org

Conversion of PRPS Hexamer to Monomer by AMPK-Mediated Phosphorylation Inhibits Nucleotide Synthesis in Response to Energy Stress Xu Qian 1 , Xinjian Li 1 , Lin Tan 2 , Jong-Ho Lee 1 , Yan Xia 1 , Qingsong Cai 1 , Yanhua Zheng 1 , Hongxia Wang 3 , Philip L. Lorenzi 2 , and Zhimin Lu 1 , 4 , 5

RESEARCH ARTICLE

ABSTRACT Tumors override energy stress to grow. However, how nucleotide synthesis is regulated under energy stress is unclear. We demonstrate here that glucose dep-

rivation or hypoxia results in the AMPK-mediated phosphorylation of phosphoribosyl pyrophosphate synthetase 1 (PRPS1) S180 and PRPS2 S183, leading to conversion of PRPS hexamers to monomers and thereby inhibiting PRPS1/2 activity, nucleotide synthesis, and nicotinamide adenine dinucleotide (NAD) production. Knock-in of nonphosphorylatable PRPS1/2 mutants, which have uninhibited activ-ity, in brain tumor cells under energy stress exhausts cellular ATP and NADPH and increases reactive oxygen species levels, thereby promoting cell apoptosis. The expression of those mutants inhibits brain tumor formation and enhances the inhibitory effect of the glycolysis inhibitor 2-deoxy- D -glucose on tumor growth. Our fi ndings highlight the signifi cance of recalibrating tumor cell metabolism by fi ne-tuning nucleotide and NAD synthesis in tumor growth.

SIGNIFICANCE: Our fi ndings elucidate an instrumental function of AMPK in direct regulation of nucleic acid and NAD synthesis in tumor cells in response to energy stress. AMPK phosphorylates PRPS1/2, converts PRPS1/2 hexamers to monomers, and inhibits PRPS1/2 activity and subsequent nucleotide and NAD synthesis to maintain tumor cell growth and survival. Cancer Discov; 8(1); 94–107. ©2017 AACR.

1 Brain Tumor Center and Department of Neuro-Oncology, The University of Texas MD Anderson Cancer Center, Houston, Texas . 2 Department of Bioin-formatics and Computational Biology, and The Proteomics and Metabo-lomics Core Facility, The University of Texas MD Anderson Cancer Center, Houston, Texas. 3 The Institute of Cell Metabolism and Disease, Shanghai Key Laboratory of Pancreatic Cancer, Shanghai General Hospital, School of Medicine, Shanghai Jiaotong University, Shanghai, China. 4 Department of Molecular and Cellular Oncology, The University of Texas MD Anderson Cancer Center, Houston, Texas. 5 Cancer Biology Program, MD Anderson Cancer Center UT Health Graduate School of Biomedical Sciences, The University of Texas, Houston, Texas.

Note: Supplementary data for this article are available at Cancer Discovery Online (http://cancerdiscovery.aacrjournals.org/). X. Qian and X. Li contributed equally to this article. Corresponding Author: Zhimin Lu , The University of Texas MD Anderson Cancer Center, 1515 Holcombe Boulevard, Unit 1002, Houston, TX 77030. Phone : 713-834-6231; Fax: 713-834-6230; E-mail: [email protected] doi: 10.1158/2159-8290.CD-17-0712 ©2017 American Association for Cancer Research.

Research. on October 13, 2020. © 2018 American Association for Cancercancerdiscovery.aacrjournals.org Downloaded from

Published OnlineFirst October 26, 2017; DOI: 10.1158/2159-8290.CD-17-0712

Page 2: Conversion of PRPS Hexamer to Monomer by AMPK-Mediated … · 94 | CANCER DISCOVERY January 2018 Conversion of PRPS Hexamer to Monomer by AMPK-Mediated Phosphorylation Inhibits Nucleotide

January 2018 CANCER DISCOVERY | 95

INTRODUCTION

Cancer metabolism is programmed to facilitate the sur-vival and proliferation of cancer cells in the nonnative tumor microenvironment (1). A rapidly growing tumor encounters nutritional stresses and recalibrates cell metabolism to survive and grow. Many tumor cells exhibit enhanced Warburg effect, which is a high rate of glycolysis followed by lactic acid fer-mentation in the cytosol even in the presence of ample oxygen (2, 3). The enhanced production of glycolytic intermediates in the tumor can be used to synthesize cellular building blocks (nucleotides, amino acids, and lipids) to meet the demands of cell proliferation. The pentose phosphate pathway (PPP), which is derived from glycolytic intermediates glucose-6-phosphate and fructose-6-phosphate, produces ribose-5-phosphate (R5P) for the de novo synthesis of nucleotides and nucleic acids (4, 5). Growth signaling through the mechanistic target of rapamycin complex 1 (mTORC1) pathway stimulates de novo pyrimidine and purine synthesis (6–8). Phosphoribosyl pyrophosphate synthetase (PRPS) catalyzes the first and rate-limiting reaction for nucleotide synthesis, producing phosphoribosyl pyrophos-phate (PRPP) from R5P by transferring the β, γ-diphosphoryl

moiety of ATP to the C1-hydroxy group of R5P (9, 10). PRPP is then used for the synthesis of purine and pyrimidine nucleo-tides (Supplementary Fig. S1A), the pyridine nucleotide cofac-tors nicotinamide adenine dinucleotide (NAD) and NADP, and the amino acids histidine and tryptophan (11).

The human PRPS family has three isoforms that share very high sequence identity: PRPS1 and PRPS2, which have 95% amino acid sequence identity, are expressed in a wide range of tissues, whereas PRPS3 is expressed specifically in the testis. PRPS1–3 are activated by Mg2+, sulfate (SO4

2−), and phos-phate, whereas PRPS1 is inhibited by the nucleotide biosyn-thesis products ADP, AMP, and GDP (12, 13). PRPS1 forms a hexamer, which is facilitated by ATP (14). The catalytic active site, which consists of the ATP binding site and the R5P binding site, is located at the interface of two domains of one subunit; the allosteric site for phosphate and ADP is located at the interfaces between three subunits of the hexamer (13), indicating that a hexamer is required for PRPS1 activity. Ketohexokinase-A (KHK-A; also known as fructokinase-A) phosphorylates PRPS1 T225 and activates PRPS1 by block-ing the binding of ADP, AMP, and GDP, which is required for hepatocellular carcinoma development (15, 16). Mutations of

Research. on October 13, 2020. © 2018 American Association for Cancercancerdiscovery.aacrjournals.org Downloaded from

Published OnlineFirst October 26, 2017; DOI: 10.1158/2159-8290.CD-17-0712

Page 3: Conversion of PRPS Hexamer to Monomer by AMPK-Mediated … · 94 | CANCER DISCOVERY January 2018 Conversion of PRPS Hexamer to Monomer by AMPK-Mediated Phosphorylation Inhibits Nucleotide

Qian et al.RESEARCH ARTICLE

96 | CANCER DISCOVERY January 2018 www.aacrjournals.org

PRPS1, which reduced the feedback inhibition of purine bio-synthesis, were identified in relapsed childhood B-cell acute lymphoblastic leukemia (ALL; ref. 17). In addition, PRPS2 was shown to be crucial for cancer cell survival (18–20). How-ever, the mechanism through which PRPS and nucleotide synthesis are regulated under energy stress is unclear.

In this study, we showed that glucose deprivation results in the AMPK-mediated phosphorylation of PRPS1 S180 and PRPS2 S183, disruption of the PRPS1/2 hexamers, and inhi-bition of PRPS1/2 activity and nucleic acid synthesis. The expression of nonphosphorylatable PRPS1/2 mutants greatly decreased cellular ATP and NADPH levels, increased reactive oxygen species (ROS) levels and cell apoptosis, and inhibited brain tumorigenesis.

RESULTSEnergy Stresses Induce Rapid Inhibition of PRPS1/2 Activity and Nucleic Acid Synthesis

To determine the effects of energy stress on the regula-tion of nucleic acid synthesis, we removed glucose from the culture medium of U87 and U251 glioblastoma (GBM) cells for 3 hours or treated the cells with the glucose metabolism inhibitor 2-deoxy-d-glucose (2-DG) for 4 hours, followed by incubation of a limited amount of D-[6-14C] glucose (0.01 mmol/L). We found that glucose deprivation (Fig. 1A) or 2-DG treatment (Supplementary Fig. S1B) largely decreased the production of glucose-derived 14C-RNA and 14C-DNA. In line with this finding, the levels of both purine (IMP, AMP, and GMP) and pyrimidine (UMP and CMP) interme-diates were decreased in U87 (Fig. 1B) and U251 cells (Sup-plementary Fig. S1C) upon glucose deprivation. However, the amount of R5P was not affected by such a short period of glucose deprivation (Fig. 1C), strongly suggesting that the decrease in nucleotide production in response to acute glucose deprivation was not regulated through PPP-derived R5P production. PRPS-catalyzed conversion of R5P to PRPP is a rate-limiting reaction (11). Quantification of PRPS1 and PRPS2 mRNA levels by PCR amplification of their cDNA, in which the PRPS1 but not the PRPS2 fragment was cut by the HindIII restriction enzyme, showed that both PRPS1 and PRPS2 were comparably expressed in U87 and U251 cells (Supplementary Fig. S1D). We immunoprecipitated PRPS1/2 with an antibody recognizing both PRPS1 and PRPS2 from U87 and U251 cells and showed that their activities were inhibited by glucose deprivation (Fig. 1D) and 2-DG treat-ment (Supplementary Fig. S1E). In line with this finding, PRPP levels were decreased by glucose deprivation (Fig. 1E) and 2-DG treatment (Supplementary Fig. S1F).

To further determine whether acute PRPS1/2 regulation is independent of glucose level and whether it is also regulated by other metabolic stress, we treated U87 cells with hypoxia for a short period (6 hours) in the presence of an ample glucose supply. Hypoxia stimulation resulted in reduced PRPS1/2 activity (Fig. 1F) and PRPP level (Fig. 1G) without affecting R5P production (Supplementary Fig. S1G). These results indicate that PRPS1/2 inhibition and the subsequent decrease of nucleic acid synthesis are rapid responses of cells to energy stresses induced by glucose deprivation and hypoxia.

AMPK Phosphorylates PRPS1 S180 and PRPS2 S183 in Response to Glucose Deprivation and Hypoxic Stimulation

Posttranslational modifications of proteins often result in the alteration of their structures and activities (21). We immu-noprecipitated FLAG-tagged PRPS1 from glucose-deprived U87 cells and performed a liquid chromatography–coupled Orbitrap tandem mass spectrometric analysis of tryptic digests of immunoprecipitated PRPS1. We found that PRPS1 was phosphorylated at S180 (Supplementary Fig. S2A), which corresponds to S183 in PRPS2 (Supplementary Fig. S2B); both residues are conserved among species (Supplementary Fig. S2C). Of note, the Scansite analyses (http://scansite3.mit.edu/) of PRPS1/2 protein sequences revealed that PRPS1 S180 and PRPS2 S183 are potential phosphorylation residues of AMP-activated protein kinase (AMPK) with −4 (Lys), −3 (Arg), −2 (Val), and +3 (Asp) residues matched to the AMPK recognition motifs (Supplementary Fig. S2D; refs. 22, 23). Glucose deprivation or 2-DG treatment, which activated AMPK, as reflected by phosphorylation of AMPKα T172 and its substrate acetyl-CoA carboxylase 1 (ACC1) S79, induced enhanced binding of endogenous AMPKα to endogenous PRPS1/2 (Fig. 2A) in U87 cells. These results were further vali-dated by coimmunoprecipitation of HA-AMPKα with FLAG-PRPS1 or FLAG-PRPS2 (Supplementary Fig. S2E).

To determine whether AMPK phosphorylates PRPS, we performed an in vitro phosphorylation analysis by incubating purified bacteria-expressed human His-PRPS1 or His-PRPS2 with active AMPK protein in the presence of [γ-32P]ATP, fol-lowed by autoradiography and immunoblot analyses by a specific anti–phospho-PRPS1/2 S180/S183 antibody, which recognizes both PRPS1 S180 and PRPS2 S183 phospho-rylation (Supplementary Fig. S2F). We found that AMPK phosphorylated PRPS1 S180 and PRPS2 S183; those phos-phorylations were abrogated by mutating the Ser residues into Ala (Fig. 2B). In addition, stoichiometry analyses of PRPS phosphorylation by AMPK showed incorporation of ∼1 mol of phosphate per mol of PRPS1 or PRPS2 protein (Supplementary Fig. S2G). These results, which are consist-ent with the finding that S180A mutation in PRPS1 and S183A mutation in PRPS2 abolished AMPK-mediated phos-phorylation (Fig. 2B), suggest that S180 and S183 are the only AMPK-phosphorylated residues in PRPS1 and PRPS2, respectively.

Consistent with the in vitro results, treatment with 2-DG (Fig. 2C) or A769662 AMPK activator (Fig. 2D) induced the phosphorylation of PRPS1 S180 and PRPS2 S183, but not PRPS1 S180A or PRPS2 S183A mutant, in U87 cells. Notably, PRPS1 S180 and PRPS2 S183 phosphorylation was rapidly induced by 2-DG treatment; this enhanced phos-phorylation was detected within 15 minutes of treatment and was sustained for at least 8 hours (Supplementary Fig. S2H). In contrast, treatment of U87 cells with compound C, a nonselective AMPK inhibitor (Fig. 2E), or deficiency of AMPKα1 and AMPKα2 in mouse embryonic fibroblasts (MEF; Fig. 2F) blocked 2-DG– (Fig. 2E), glucose-deprivation–, and A769662- (Fig. 2F) induced PRPS1/2 S180/S183 phos-phorylation. Furthermore, the inhibitory effect of compound C on 2-DG–induced PRPS1/2 S180/S183 phosphorylation

Research. on October 13, 2020. © 2018 American Association for Cancercancerdiscovery.aacrjournals.org Downloaded from

Published OnlineFirst October 26, 2017; DOI: 10.1158/2159-8290.CD-17-0712

Page 4: Conversion of PRPS Hexamer to Monomer by AMPK-Mediated … · 94 | CANCER DISCOVERY January 2018 Conversion of PRPS Hexamer to Monomer by AMPK-Mediated Phosphorylation Inhibits Nucleotide

AMPK Phosphorylates PRPS to Reduce Nucleotide Synthesis RESEARCH ARTICLE

January 2018 CANCER DISCOVERY | 97

was also detected in U251 cells, LKB1-overexpressed A549 (A549/LKB1) human non–small cell lung cancer cells, BxPC-3 human pancreatic cancer cells, and MDA-MB-231 human breast cancer cells (Supplementary Fig. S2I). As expected, hypoxia also induced PRPS1/2 S180/S183 phosphorylation, and compound C treatment inhibited this phosphorylation (Fig. 2G). These results indicate that AMPK activation, which is induced by glucose deprivation and hypoxic stimulation, is sufficient to phosphorylate PRPS1 S180 and PRPS2 S183.

AMPK-Mediated PRPS1/2 Phosphorylation Converts PRPS1/2 Hexamers to Monomers

Structural analyses of PRPS1 hexamer revealed that PRPS1 S180 is located proximal to the interface of the subunits of the hexamer (Fig. 3A). We hypothesized that such phos-phorylation regulates the hexameric structure of PRPS1/2. To test that, we immunoprecipitated endogenous PRPS1/2 and performed size-exclusion chromatography. According to the molecular weight calibration standard (Supplementary

Figure 1.  Energy stresses induce rapid inhibition of PRPS1/2 and suppress nucleic acid synthesis. A, C, D, and F, Data represent the mean ± SD from three independent experiments. B, E, and G, Data represent the mean ± SD from triplicate samples. *, P < 0.05; **, P < 0.001; n.s., no significance. A, U87 or U251 cells were cultured in the presence or absence of glucose (Glc) for 3 hours, followed by labeling with D-[6-14C] glucose (1 μCi, 0.01 mmol/L) for 0.5 hours. The amounts of 14C-RNA and 14C-DNA were measured. B, U87 cells were cultured in the presence or absence of Glc for 3 hours. Levels of nucleotide monophosphate were measured by LC/MS-MS. C, U87 or U251 cells were cultured in the presence or absence of Glc for 3 hours. R5P levels were measured. D and E, U87 or U251 cells were cultured in the presence or absence of Glc for 3 hours. The activity of immunoprecipitated PRPS1/2 was measured. Immunoblot analysis was performed with the indicated antibody (D). Levels of PRPP were measured by LC/MS-MS (E). F and G, U87 or U251 cells were cultured with or without hypoxia for 6 hours. The activity of immunoprecipitated PRPS1/2 was measured. Immunoblot analysis was performed with the indicated antibody (F). Levels of PRPP were measured by LC/MS-MS (G).

A

B

E F G

C D

Glc +

+

+ − + − − + − +

− + − + − + − + −

Glc −Glc +Glc −

Glc +Glc −

** ** ** ** n.s.

Glc +Glc −

** **

n.s.

1.2 1.2

1.0

0.8

0.6

0.4

0.2

0

1.2

1.2

1.0

0.8

0.6

0.4

0.2

0U87

IP: PRPS1/2, WB: PRPS1/2

IP: PRPS1/2, WB: PRPS1/2

U251

1.0

0.8

0.6

0.4

0.2

0

1.0

0.8

0.6

Rel

ativ

e ra

diol

abel

edR

NA

/tota

l RN

A

Rel

ativ

e ra

diol

abel

edD

NA

/tota

l DN

A

Rel

ativ

e R

5P le

vel

Rel

ativ

e P

RP

S a

ctiv

ity

0.4

0.2

0

IMP AMP GMP UMP CMP1.4 70 3.5 50 1.2

1.0

0.8

0.6

0.4

0.2

0

40

30

20

10

0

3.0

2.5

2.0

1.5

1.0

0.5

0

60

50

40

30

20

10

0

1.2

*

* *

* *

**

**

**

1.0

0.8

0.6

Pea

k ar

ea (×

106 )

0.4

0.2

0

4.2 30

1.2

5 35

30

25

20

15

10

5

0

4

3

Pea

k ar

ea (×

106 )

of P

RP

P

2

1

0HypoxiaIgG

PRPS1/2

1.0

0.8

0.6

0.4

0.2

0U87 U251

25

20

15

10

5

0

Pea

k ar

ea (×

106 )

of P

RP

P

Rel

ativ

e P

RP

S a

ctiv

ity

3.5

2.8

2.1

1.4

0.7

0Glc

Glc

U251 U87U87

U87 U251 U87 U251

Hypoxia −Hypoxia +

U251 U87 U251lgG

PRPS1/2

Research. on October 13, 2020. © 2018 American Association for Cancercancerdiscovery.aacrjournals.org Downloaded from

Published OnlineFirst October 26, 2017; DOI: 10.1158/2159-8290.CD-17-0712

Page 5: Conversion of PRPS Hexamer to Monomer by AMPK-Mediated … · 94 | CANCER DISCOVERY January 2018 Conversion of PRPS Hexamer to Monomer by AMPK-Mediated Phosphorylation Inhibits Nucleotide

Qian et al.RESEARCH ARTICLE

98 | CANCER DISCOVERY January 2018 www.aacrjournals.org

Figure 2.  AMPK phosphorylates PRPS1 S180 and PRPS2 S183 in response to glucose (Glc) deprivation and hypoxic stimulation. Immunoprecipitation or immunoblot analyses were performed with the indicated antibodies. A, U87 cells were cultured in the presence or absence of Glc for 3 hours or treated with or without 2-DG (25 mmol/L) for 4 hours. B, Bacterially purified WT His-PRPS1 or His-PRPS1 S180A mutant (top) or WT His-PRPS2 or His-PRPS2 S183A mutant (bottom) was incubated with or without bacterially purified active AMPK in the presence of [γ-32P]ATP. Autoradiography and immunoblot analyses were performed. C and D, U87 cells expressing WT FLAG-PRPS1 or FLAG-PRPS1 S180A (top) or WT FLAG-PRPS2 or FLAG-PRPS2 S183A (bottom) were treated with or without 2-DG (25 mmol/L; C) or A769662 (0.5 mmol/L; D) for 4 hours. E U87 cells were pretreated with or without compound C (5 μmol/L) for 30 minutes and then treated with or without 2-DG (25 mmol/L) for 4 hours. F, WT or AMPKα1/2 double-knockout (DKO) MEFs were cultured in the pres-ence or absence of Glc for 3 hours (left) or treated with or without A769662 (0.5 mmol/L, right) for 4 hours. G, U87 cells were pretreated with or without compound C (5 μmol/L) for 30 minutes and cultured in hypoxia for the indicated time periods.

GIc

A

E F G

B C D

WB: PRPS1/2 His-PRPS1 WT

32P-PRPS1

32P-PRPS2

+− +−S180A

AMPK

FLAG-PRPS1 FLAG-PRPS1

A769662

IP: FLAG IP: FLAG

IP: FLAG IP: FLAG

WT− + − +

S180A WT− + − +

S180A2-DG

WT− + − +

S183A2-DG

FLAG-PRPS2FLAG-PRPS2 WT− + − +

S183AA769662

His-PRPS2 WT− + − +

S183A

WT+ − + −

DKO WT+− +− − − − − +

0 4 8 12 4

DKO

AMPK

IP: PRPS1/2

Input

WB: PRPS1/2

WB: AMPKα

WB: AMPKα WB: α-TubulinWB: α-Tubulin

WB: α-Tubulin

WB: AMPKα

WB: AMPKα WB: AMPKα

AMPKα1/2Glc

AMPKα1/2

Compound CHypoxia (h)

A769662

WB: PRPS1/2 pS180/S183

WB: PRPS1/2 pS180/S183 WB: PRPS1/2 pS180/S183

WB: PRPS1/2 pS180/S183WB: PRPS1/2 pS180/S183

WB: PRPS1/2 pS180/S183

WB: PRPS1/2 pS180/S183WB: PRPS1/2 pS180/S183 WB: PRPS1/2 pS180/S183

WB: PRPS1/2 pS180/S183

WB: PRPS1/2WB: PRPS1/2 WB: PRPS1/2

WB: PRPS1/2

WB: PRPS1

WB: FLAG WB: FLAG

WB: FLAGWB: FLAG

WB: ACC1 pS79 WB: ACC1 pS79

WB: ACC1 pS79WB: ACC1 pS79

WB: ACC1 pS79WB: ACC1 pS79 WB: ACC1 pS79

WB: ACC1 pS79

WB: ACC1 WB: ACC1

WB: ACC1WB: ACC1

WB: ACC1WB: ACC1 WB: ACC1

WB: ACC1

WB: PRPS2

WB: AMPKα pT172

WB: AMPKα

WB: AMPKα

WB: α-Tubulin

WB: α-Tubulin

WB: α-Tubulin WB: α-TubulinWB: α-Tubulin

Compound C2-DG

WB: α-Tubulin

WB: ACC1 pS79

WB: ACC1

WB: PRPS1/2

IP: AMPKαAMPKαIgG

IgG+−

+− −

− ++

− − +− ++

2-DG

IgG

PRPS1/2

Fig. S3A), the fractions 17 to 19 that coeluted with a molecu-lar weight around 200 kDa contained hexamers of PRPS1/2, which were shown to be nonphosphorylated (Fig. 3B, top). In contrast, the fractions 42 to 44 that coeluted with a molecular weight around 29 kDa contained monomers of PRPS1/2 (34 kDa), which were phosphorylated (Fig. 3B, bot-tom). Incubation of wild-type (WT) or nonphosphorylatable mutants of PRPS1/2 in the presence or absence of active AMPK revealed that both WT and mutated PRPS1/2 existed as hexamers in the absence of AMPK (Fig. 3C). However,

the presence of AMPK, which phosphorylated WT but not mutated PRPS1/2, converted WT but not mutated PRPS1/2 into a monomer.

To further validate our findings in vivo, we used CRISPR/Cas9 genome editing knock-in technology to simultaneously replace endogenous PRPS1/2 with PRPS1 S180A and PRPS2 S183A (PRPS1/2-2S/A, Supplementary Fig. S3B–S3D) in U87 cells (24). Glucose deprivation (Fig. 3D) or A769662 treat-ment (Supplementary Fig. S3E) converted WT PRPS1/2 but not their mutants from hexamers into monomers. These

Research. on October 13, 2020. © 2018 American Association for Cancercancerdiscovery.aacrjournals.org Downloaded from

Published OnlineFirst October 26, 2017; DOI: 10.1158/2159-8290.CD-17-0712

Page 6: Conversion of PRPS Hexamer to Monomer by AMPK-Mediated … · 94 | CANCER DISCOVERY January 2018 Conversion of PRPS Hexamer to Monomer by AMPK-Mediated Phosphorylation Inhibits Nucleotide

AMPK Phosphorylates PRPS to Reduce Nucleotide Synthesis RESEARCH ARTICLE

January 2018 CANCER DISCOVERY | 99

results indicate that AMPK-mediated PRPS1/2 phosphoryla-tion disrupts PRPS1/2 hexamers and converts them into monomers.

AMPK-Mediated PRPS1/2 Phosphorylation Inhibits PRPS1/2 Activity, Nucleotide Synthesis, and NAD Production

PRPS1 may exist as a hexamer to maintain PRPS1 activity (13). To determine the effect of AMPK-dependent PRPS1/2 phosphorylation on the regulation of PRPS1/2 activity, we performed an in vitro phosphorylation assay by mixing WT or nonphosphorylatable mutants of PRPS1/2, with or without active AMPK. The presence of AMPK largely inhib-ited the activity of WT PRPS1 and WT PRPS2 but not of their mutants (Fig. 4A). Consistent with those results, glu-cose deprivation largely inhibited the activity of ectopically expressed WT FLAG-PRPS1 and FLAG-PRPS2 (Fig. 4B) and endogenously expressed PRPS1/2 (Fig. 4C). Similar inhibi-tion was also observed by treating U87 cells with A769662 AMPK activator (Supplementary Fig. S4A and S4B). In contrast, glucose deprivation or A769662 treatment did not affect the activity of ectopically expressed FLAG-PRPS1

S180A or FLAG-PRPS2 S183A (Fig. 4B; Supplementary Fig. S4A) or knock-in–expressed PRPS1 S180A and PRPS2 S183A (Fig. 4C; Supplementary Fig. S4B). These results indicate that glucose deprivation or activation of AMPK inhibits PRPS1/2 activity by the AMPK-mediated phospho-rylation of PRPS1/2 and subsequent disruption of PRPS1/2 hexamers.

The allosteric site of PRPS1, which binds phosphate for activation or ADP and AMP for inhibition, is located at the interfaces among three subunits of the hexamer (12, 13). We found that purified WT PRPS1 and PRPS1 S180A mutant were activated by phosphate or inhibited by ADP and AMP to a similar extent (Supplementary Fig. S4C), suggesting that mutation of this phosphorylation residue does not affect allosteric regulation of PRPS1. In contrast, the presence of AMPK, which phosphorylated and converted WT PRPS1 hex-amers to monomers, made PRPS1 unresponsive to regulation by phosphate, ADP, or AMP (Supplementary Fig. S4C). These results suggest that rapid PRPS1 phosphorylation by AMPK disrupted the allosteric regulation of PRPS1 by converting PRPS1 hexamers to monomers and is an early response of tumor cells to energy stress.

Figure 3.  AMPK-mediated PRPS1/2 phosphorylation converts PRPS1/2 hexamers to monomers. B–D, Immunoprecipitation and immunoblot analyses were performed using the indicated antibodies. A, Structural view of PRPS1 (PDB: 2H06). Three homodimers formed a hexamer in a propeller shape, in which the N-terminal domains formed the inner circle and the C-terminal domains formed the “blades” toward the outside. Serine 180 (in yellow) was located close to the intersurface between each homodimer. B, U87 cell lysates were fractionated on a size-exclusion column. PRPS1/2 was eluted in frac-tion F17-19 (around 200 kDa) or in F42-44 (around 29 kDa) on the same column under identical conditions, as determined by chromatography molecular weight calibration. C, Purified WT FLAG-PRPS1 or FLAG-PRPS1 S180A (left) or WT FLAG-PRPS2 or FLAG-PRPS2 S183A (right) from U87 cells was incubated with or without active AMPK in the presence of ATP (0.5 mmol/L). Fractionation was performed on a size-exclusion column. The indicated frac-tions were used for the immunoblot analyses. D, Parental U87 cells or U87 cells with knock-in of PRPS1/2 S180A/S183A (PRPS1/2-2S/A) were cultured in the presence or absence of glucose for 3 hours. Cell lysates were fractionated on a size-exclusion column. The indicated fractions were used for the immunoblot analyses.

A

90°

FLAG-P

RPS1

AMPK

+

+

180°

200k

200k 150k

WT

S18

0A

++

+

−−

+

WT

PR

PS

1/2-

2S/A

Par

enta

l

S18

3A

66k 29k

FLAG-P

RPS2

AMPK

Glu

cose

200k 150k 66k 29k

200k 150k 66k 29k

66k 29k

F 16

F

F 17 18 19 24 25 33

WB: PRPS1/2 pS180/S183

WB: PRPS1/2 pS180/S183

WB: PRPS1/2 pS180/S183

WB: PRPS1/2 pS180/S183

WB: FLAG

WB: FLAG

WB: FLAG

WB: FLAG

WB: PRPS1/2 pS180/S183

WB: PRPS1/2

WB: PRPS1/2

WB: PRPS1/2

WB: PRPS1/2

WB: PRPS1/2 pS180/S183

WB: PRPS1/2 pS180/S183

WB: PRPS1/2 pS180/S183

WB: FLAG

WB: FLAG

WB: FLAG

WB: FLAG

34 42 43 44 F 17 18 19 24 25 33 34 42 43 44

F 17 18 19 24 25 33 34 42 43 44

31 32 33 34 35 36 37 38 39 40 41 42 43 44 45

17 18 19 20 21 22 23

WB: PRPS1/2 pS180/S183

WB: PRPS1/2

WB: PRPS1/2 pS180/S183

WB: PRPS1/2

24 25 26 27 28 29 30

150k

C

D

B

Research. on October 13, 2020. © 2018 American Association for Cancercancerdiscovery.aacrjournals.org Downloaded from

Published OnlineFirst October 26, 2017; DOI: 10.1158/2159-8290.CD-17-0712

Page 7: Conversion of PRPS Hexamer to Monomer by AMPK-Mediated … · 94 | CANCER DISCOVERY January 2018 Conversion of PRPS Hexamer to Monomer by AMPK-Mediated Phosphorylation Inhibits Nucleotide

Qian et al.RESEARCH ARTICLE

100 | CANCER DISCOVERY January 2018 www.aacrjournals.org

Figure 4.  AMPK-mediated PRPS1/2 phosphorylation inhibits PRPS1/2 activity and nucleotide synthesis. A–C, Immunoprecipitation and immunoblot analyses were performed using the indicated antibodies. A–C, G, and H, Data represent the mean ± SD from three independent experiments. D–F, and I, Data represent the mean ± SD from triplicate samples. *, P < 0.05; **, P < 0.001; n.s., no significance. A, Purified WT FLAG-PRPS1 or FLAG-PRPS1 S180A (left) or WT FLAG-PRPS2 or FLAG-PRPS2 S183A (right) from U87 cells was incubated with or without active AMPK in the presence of ATP (0.5 mmol/L). After the removal of AMPK by Ni-NTA resin, the activities of the remaining PRPS1/2 proteins were measured. B, U87 cells expressing WT FLAG-PRPS1, FLAG-PRPS1 S180A (left), WT FLAG-PRPS2, or FLAG-PRPS2 S183A (right) were cultured in the presence or absence of glucose (Glc) for 3 hours. The activity of immunoprecipitated FLAG-PRPS1/2 was measured. C, Parental U87 cells or U87 cells with knock-in of PRPS1/2-2S/A were cultured in the presence or absence of Glc for 3 hours. The activity of immunoprecipitated PRPS1/2 was measured. PRPS1/2-2S/A-C1, clone 1; PRPS1/2-2S/A-C2, clone 2. D, Parental U87 cells or U87 cells with knock-in of PRPS1/2-2S/A were pretreated with or without A769662 (0.5 mmol/L) for 3 hours, followed by metabolic labeling with 13C6-glucose (10 mmol/L) for 30 minutes. 13C-labeled metabolite intermediates of PRPP, IMP, AMP, GMP, UMP, and CMP were measured by LC/MS-MS. E and F, Parental U87 cells or U87 cells with knock-in of PRPS1/2-2S/A were pretreated with or without A769662 (0.5 mmol/L) for 3 hours, followed by metabolic labeling with [13C2,15N]-glycine (0.4 mmol/L; E) or [13C5,15N4]-hypoxanthine (0.1 mmol/L; F) for 2 hours. 13C-labeled IMP was measured by LC/MS-MS. G and H, Parental U87 cells or U87 cells with knock-in of PRPS1/2-2S/A were pretreated with or without A769662 (0.5 mmol/L) for 6 hours (G) or hypoxia for 12 hours (H), followed by labeling with D-[6-14C] glucose (1 μCi, 0.01 mmol/L) for 30 minutes. The amounts of 14C-RNA (left) and 14C-DNA (right) were measured. I, Parental U87 cells or U87 cells with knock-in of PRPS1/2-2S/A were pretreated with or without A769662 (0.5 mmol/L) for 3 hours, followed by metabolic labeling with 13C6-glucose (10 mmol/L) for 30 minutes. 13C-labeled NAD was measured by LC/MS-MS.

1.2

AMPK −AMPK +

WB: FLAG

** n.s.

* n.s.

FLAG-PRPS1 WT S180A

Rel

ativ

e P

RP

S a

ctiv

ityA

D

G

H I

F

B C

E

1.0

0.8

0.6

0.4

0.2

0

12

A769662

Pea

k ar

ea (

×10

) 10

8

6

4

2

0

1.2

AMPK −AMPK +** n.s.

FLAG-PRPS2 WT S183AR

elat

ive

PR

PS

act

ivity

1.0

0.8

0.6

0.4

0.2

0

1.2

Glc +Glc −

** n.s.

** n.s.

* n.s.

n.s.

FLAG-PRPS1 WT S180A

Rel

ativ

e P

RP

S a

ctiv

ity

1.0

0.8

0.6

0.4

0.2

0

1.2

Glc +Glc −

Glc +Glc −

IgG

* n.s.

FLAG-PRPS2 WT S183A

Rel

ativ

e P

RP

S a

ctiv

ity

Rel

ativ

e P

RP

S a

ctiv

ity

Paren

tal

PRPS1/2

-2S/A

-C1

PRPS1/2

-2S/A

-C2

1.0

0.8

0.6

0.4

0.2

0

1.2

1.0

0.8

0.6

0.4

0.2

0

** n.s.

A769662 −A769662 +

n.s.

Rel

ativ

e ra

diol

abel

edR

NA

/tota

l RN

A

Paren

tal

PRPS1/2

-2S/A

-C1

PRPS1/2

-2S/A

-C2

1.2

1.0

0.8

0.6

0.4

0.2

0

0.6

0.5

0.4

0.3

0.2

0.1

0A769662

Parental PRPS1/2-2S/A-C1

Parental PRPS1/2-2S/A-C1

− +

− + − +

*18

15

12

9

6

3

0

Parental PRPS1/2-2S/A-C1

− + − +

* n.s.

100

0

20

40

60

80

Parental PRPS1/2-2S/A-C1

− + − +

* n.s. * n.s. * n.s.

14 8

0 0

2

4

6

4

0

1

2

3

2

4

6

8

10

12

Parental PRPS1/2-2S/A-C1

− + − +Parental PRPS1/2-

2S/A-C1Parental PRPS1/2-

2S/A-C1

− + − + − + − +

− +

* n.s.

9

78

56

4

23

10

A769662Parental PRPS1/2-

2S/A-C1

− + − +

A769662Parental PRPS1/2-

2S/A-C1

− + − +

** n.s.

A769662 −A769662 +

n.s.

Rel

ativ

e ra

diol

abel

edD

NA

/tota

l DN

A

Paren

tal

PRPS1/2

-2S/A

-C1

PRPS1/2

-2S/A

-C2

1.2

1.0

0.8

0.6

0.4

0.2

0

* n.s.

Hypoxia −Hypoxia +

n.s.

Rel

ativ

e ra

diol

abel

edR

NA

/tota

l RN

A

Paren

tal

PRPS1/2

-2S/A

-C1

PRPS1/2

-2S/A

-C2

1.2

1.0

0.8

0.6

0.4

0.2

0

* n.s.

Hypoxia −Hypoxia +

n.s.

Rel

ativ

e ra

diol

abel

edD

NA

/tota

l DN

A

Paren

tal

PRPS1/2

-2S/A

-C1

PRPS1/2

-2S/A

-C2

1.2

1.0

0.8

0.6

0.4

0.2

0

*14

12

10

8

6

4

2

0

PRPS1/2

IP: PRPS1/2, WB: PRPS1/2

C-NAD

C-PRPP C-IMP C-AMP C-GMP C-UMP C-CMP

WB: FLAG WB: FLAG WB: FLAG

Research. on October 13, 2020. © 2018 American Association for Cancercancerdiscovery.aacrjournals.org Downloaded from

Published OnlineFirst October 26, 2017; DOI: 10.1158/2159-8290.CD-17-0712

Page 8: Conversion of PRPS Hexamer to Monomer by AMPK-Mediated … · 94 | CANCER DISCOVERY January 2018 Conversion of PRPS Hexamer to Monomer by AMPK-Mediated Phosphorylation Inhibits Nucleotide

AMPK Phosphorylates PRPS to Reduce Nucleotide Synthesis RESEARCH ARTICLE

January 2018 CANCER DISCOVERY | 101

PRPS1 S103T and A190T mutations, identified in ALL, reduce feedback inhibition by nucleotide biosynthesis prod-ucts (17). S103 is in the allosteric site of PRPS1 hexamer, and A190 is in the dimer interface. It was previously shown that mutations of these two residues destabilize the dimer inter-face and allosteric sites for ADP or GDP (17). We treated U87 cells expressing FLAG-PRPS1 S103T or FLAG-PRPS1 A190T mutant with A769662 and found that AMPK activation induced S180 phosphorylation of both mutants (Supple-mentary Fig. S4D). Size-exclusion chromatography showed that the presence of AMPK converted both mutants from hexamers to monomers (Supplementary Fig. S4E). These data suggest that AMPK-mediated PRPS1 phosphorylation is able to inhibit PRPS1 S103T and PRPS1 A190T mutants identified in ALL.

To determine whether PRPS1/2 inhibition by acti-vated AMPK decreases nucleotide synthesis, we used tar-geted LC/MS-MS to measure the incorporation of stable isotope-labeled glucose (13C6-glucose) into purine and pyrimidine intermediates. A769662 treatment of U87 cells significantly decreased the amount of 13C-labeled PRPP, and purine (IMP, AMP, and GMP) and pyrimidine (UMP and CMP) intermediates; that decrease was abro-gated by knock-in expression of PRPS1 S180A and PRPS2 S183A (Fig. 4D). To determine whether AMPK-meditated PRPS1/2 inhibition decreases nucleotide synthesis through de novo or salvage pathways, we used [13C2,15N]-glycine and [13C5,15N4]-hypoxanthine to label U87 cells, respectively (Supplementary Fig. S4F). A769662 treatment reduced IMP incorporation of both [13C2,15N]-glycine (Fig. 4E) and [13C5,15N4]-hypoxanthine (Fig. 4F); this reduction was abol-ished by knock-in expression of PRPS1 S180A and PRPS2 S183A, suggesting that both de novo and salvage nucleo-tide synthesis pathways are regulated by AMPK-mediated PRPS1/2 inhibition.

A769662 treatment (Supplementary Fig. S4G) and the expression of active AMPKα11–312 mutant (25) but not AMPKα11–312 T172A kinase-inactive mutant (Supplementary Fig. S4H) in U87 or U251 cells inhibited the synthesis of RNA and DNA. A similar inhibitory effect was observed in A769662-treated A549/LKB1, BxPC-3, and MDA-MB-231 cells (Supplementary Fig. S4I). Those inhibitory effects induced by AMPK activation were abrogated by a deficiency of AMPKα1/2 (Supplementary Fig. S4J and S4K) and com-bined knock-in expression of PRPS1 S180A and PRPS2 S183A (Fig. 4G). In addition, knock-in of PRPS1 S180A and PRPS2 S183A blocked the hypoxia-induced inhibition of the synthesis of RNA and DNA (Fig. 4H). These results indicate that activation of AMPK, induced by glucose deprivation or hypoxic stimulation, inhibits nucleic acid synthesis by AMPK-mediated phosphorylation of PRPS1/2.

In addition to being used for nucleotide synthesis, PRPP is also an intermediate for NAD synthesis (26). Glucose dep-rivation reduced levels of NAD in both U87 and U251 cells (Supplementary Fig. S4L). In addition, A769662 treatment of U87 cells cultured with 13C6-glucose reduced the amount of 13C-labeled NAD, and this reduction was abrogated by knock-in expression of PRPS1 S180A and PRPS2 S183A (Fig. 4I). These results indicate that AMPK-mediated PRPS1/2 phosphorylation inhibits NAD biosynthesis.

AMPK-Mediated PRPS1/2 Phosphorylation Copes with Energy Stress in Tumorigenesis

Nucleotide synthesis is an energy-intensive process that con-sumes several sources of carbon and nitrogen and a large amount of ATP (27). A collapse of cellular ATP levels induces apoptosis and necrosis (27–30). To determine the effect of AMPK-inhibited PRPS1/2 on cellular levels of ATP, we cul-tured U87 cells with a low amount of glucose (0.5 mmol/L) to maintain a low level of glycolysis, which resulted in partially inhibited RNA and DNA synthesis (Supplementary Fig. S5A), as evidenced by phosphorylation of PRPS1 S180 and PRPS2 S183 (Supplementary Fig. S5B). However, knock-in expression of PRPS1 S180A and PRPS2 S183A in U87 cells had no inhibi-tory effect on the nucleic acid synthesis (Supplementary Fig. S5A) and resulted in largely decreased cellular ATP levels (Fig. 5A). Similar results were obtained by treating U87 cells with hypoxia and revealed that knock-in of PRPS1 S180A and PRPS2 S183A decreased cellular ATP levels (Supplementary Fig. S5C).

NADPH eliminates ROS. The cellular level of NADPH is the balance between NADPH synthesis, which occurs mainly through the PPP and mitochondrial metabolism, and NADPH consumption, which fuels fatty-acid synthesis, redox reactions, and de novo DNA synthesis (31–33; Sup-plementary Fig. S5D). Ribonucleotide reductase (RR), which is downstream from PRPS in the de novo DNA synthesis pathway, consumes a large amount of NADPH and catalyzes the conversion of ribonucleoside diphosphates to their cor-responding deoxyribonucleotides (34, 35). Consistent with a previous report showing that AMPK maintains NADPH levels in the absence of glucose (31), we found that NADPH levels were not dramatically altered in U87 cells during the short period of limited glucose supply (Fig. 5B). However, significant decrease of NADPH levels (Fig. 5B) with large increase of ROS levels (Fig. 5C) was observed in U87 cells with knock-in of PRPS1 S180A and PRPS2 S183A. These results suggest that the unrestricted and activated PRPS-dependent nucleic acid synthesis mediated by PRPS1 S180A and PRPS2 S183A leads to consumption of a substantial amount of ATP and NADPH, leading to increased ROS levels under these metabolic stresses. As expected, knock-in of PRPS1 S180A and PRPS2 S183A enhanced apoptosis (Fig. 5D) and caspase 3–mediated PARP cleavage (Fig. 5E).

To determine the effect of AMPK-mediated PRPS1/2 activity inhibition in cell survival, we cultured cells with a low amount of glucose (0.5 mmol/L). Knock-in of PRPS1 S180A and PRPS2 S183A increased sensitivity of cells to glucose restric-tion and reduced the survival of these cells (Fig. 5F). Treating with N-acetyl-L-cysteine (NAC), a scavenger of free radicals, increased survival of U87 cells with knock-in of PRPS1 S180A and PRPS2 S183A in a short exposure time of low glucose (24 hours), but failed to do so in a long exposure time (48 hours; Fig. 5G). Consistent with the finding that AMPK-mediated PRPS1 phosphorylation inhibits PRPS1 S103T and PRPS1 A190T activity, the cell viability of Hep3B or Huh7 hepatocel-lular carcinoma cells with depletion of primarily expressed endogenous PRPS1 (16) and reconstituted expression of WT FLAG-rPRPS1, FLAG-rPRPS1 S103T, or FLAG-rPRPS1 A190T was comparable with each other in response to low glucose (0.5 mmol/L) stimulation (Supplementary Fig. S5E and S5F).

Research. on October 13, 2020. © 2018 American Association for Cancercancerdiscovery.aacrjournals.org Downloaded from

Published OnlineFirst October 26, 2017; DOI: 10.1158/2159-8290.CD-17-0712

Page 9: Conversion of PRPS Hexamer to Monomer by AMPK-Mediated … · 94 | CANCER DISCOVERY January 2018 Conversion of PRPS Hexamer to Monomer by AMPK-Mediated Phosphorylation Inhibits Nucleotide

Qian et al.RESEARCH ARTICLE

102 | CANCER DISCOVERY January 2018 www.aacrjournals.org

These results suggest that PRPS1 S103T or A190T mutation does not alter the sensitivity of cells to energy stress.

We next compared the viability of normal and tumor cells by culturing normal human astrocytes (NHA), U87 cells, U251 cells (Supplementary Fig. S5G), human pancreatic duct epithelial (HPDE) cells, and BxPC-3 pancreatic cancer cells (Supplementary Fig. S5H) in medium with a low glucose concentration. We found that tumor cells had more cell death than did normal cells. Combined treatment of compound C with a low glucose concentration caused more death of tumor cells than of normal cells. These results suggest that normal cells, which usually have limited and controlled proliferation, do not need high PRPS activity to produce nucleotides and that AMPK-dependent inhibition of PRPS activity is a pri-marily regulatory mechanism for tumor cells under energy stress that has been induced by their rapid proliferation.

To determine the role of PRPS1/2 regulation by AMPK in tumorigenesis, we intracranially injected athymic nude mice with U87 cells, with or without knock-in of PRPS1 S180A and PRPS2 S183A mutants. The expression of the PRPS1/2 mutants inhibited tumor growth (Fig. 6A and B) with abro-gated PRPS1 S180 and PRPS2 S183 phosphorylation (Fig. 6C), as detected using a specificity-validated antibody (Sup-plementary Fig. S5I), and increased tumor cell apoptosis (Fig. 6D), supporting the critical role of AMPK-meditated phosphorylation and inhibition of PRPS1/2 in counteract-ing nutritional stress to maintain tumor growth. To further

support this finding, we inhibited glucose use by tumor cells via the intraperitoneal injection of 2-DG. As expected, treat-ment with 2-DG decreased the growth of tumors expressing WT PRPS1/2 (Fig. 6A and B) and increased PRPS1 S180 and PRPS2 S183 phosphorylation (Fig. 6C). Of note, we observed that 2-DG treatment induced greater inhibition of the growth of tumors and apoptosis with knock-in of PRPS1 S180A and PRPS2 S183A mutants (Fig. 6A, B, and D). These results indi-cate that inhibition of glycolysis results in PRPS1/2 phospho-rylation and inhibition by AMPK, leading to recalibration of cellular homeostasis during tumor growth.

DISCUSSIONTumor cells reprogram their metabolism to counteract

energy stress (1). In this study, we demonstrated that glu-cose deprivation and hypoxia resulted in AMPK-mediated phosphorylation of PRPS1 S180 and PRPS2 S183. That phosphorylation converted PRPS1/2 hexamers to monomers and inhibited PRPS1/2 activity and subsequent synthesis of nucleotide and NAD (Fig. 6E).

Expression of nonphosphorylatable mutants PRPS1 S180A and PRPS2 S183A, whose activity was unaffected under low glucose condition or hypoxia condition, greatly decreased cellular ATP levels and cell survival. Those results strongly suggest that tumor cells cope with energy stress by decreas-ing nucleotide synthesis, cell proliferation, and unnecessary

Parental

A

E F G

B C D

1.2

1.0

0.8

0.6

Rel

ativ

e AT

P le

vel

NA

DP

/NA

DP

H

Via

ble

cells

(%

)

Via

ble

cells

(%

)R

elat

ive

RO

S le

vel

Apo

ptot

ic c

ells

(%

)

0.4

0.2

0

Paren

tal

PRPS1/2-

2S/A

-C1

PRPS1/2-

2S/A

-C2

0

00 24

0.5 mmol/L Glc (h)0.5 mmol/L Glc (h)

480 24 48 0

NAC−NAC+

24

0.5 mmol/L Glc (h)

48

20

40

60

80

100 100* *n.s. n.s.

80

60

40

20

0

Glc (mmol/L) 25 0.5 0.5

WB: Cleaved PARP

WB: α-Tubulin

0.525 25

Paren

tal

PRPS1/2

-2S/A

-C2

PRPS1/2

-2S/A

-C1

5

10

15

20 10

8

6

4

2

0 0

10

20

30

40

50

60** **

Glc 25 mmol/LGlc 0.5 mmol/L

Paren

tal

PRPS1/2

-2S/A

-C1

PRPS1/2

-2S/A

-C2

Paren

tal

PRPS1/2

-2S/A

-C1

PRPS1/2

-2S/A

-C2

0 h4 h8 h12 h

0 h4 h8 h12 h

* * * *

0 4 8

0.5 mmol/L Glc (h)

12

PRPS1/2-2S/A-C1PRPS1/2-2S/A-C2

Parental

PRPS1/2-2S/A-C1 PRPS1/2-2S/A-C1 PRPS1/2-2S/A-C2

PRPS1/2-2S/A-C2

Figure 5.  AMPK-mediated PRPS1/2 phosphorylation maintains tumor cell homeostasis under energy stress. A–D and F and G, Data represent the mean ± SD from three independent experiments. *, P < 0.05; **, P < 0.001; n.s., no significance. A, Parental U87 cells or U87 cells with knock-in of PRPS1/2-2S/A were cultured in 0.5 mmol/L glucose (Glc) for the indicated periods of time. Intracellular ATP levels were measured. B and C, Parental U87 cells or U87 cells with knock-in of PRPS1/2-2S/A were cultured in 0.5 mmol/L Glc for the indicated periods of time. Intracellular NADP+ and NADPH levels (B) or ROS levels (C) were measured. D and E, Parental U87 cells or U87 cells with knock-in of PRPS1/2-2S/A were cultured in 25 mmol/L or 0.5 mmol/L Glc for 8 hours. An apoptotic assay was performed (D). Immunoblot analyses were performed with the indicated antibodies (E). F, Parental U87 cells or U87 cells with knock-in of PRPS1/2-2S/A were cultured in 0.5 mmol/L Glc for the indicated periods of time. Viable cells were measured. G, U87 cells with knock-in of PRPS1/2-2S/A, pretreated with or without NAC (2 mmol/L) for 30 minutes, were cultured in 0.5 mmol/L Glc for the indicated periods of time. Viable cells were measured. Left, clone 1 (C1); right, clone 2 (C2).

Research. on October 13, 2020. © 2018 American Association for Cancercancerdiscovery.aacrjournals.org Downloaded from

Published OnlineFirst October 26, 2017; DOI: 10.1158/2159-8290.CD-17-0712

Page 10: Conversion of PRPS Hexamer to Monomer by AMPK-Mediated … · 94 | CANCER DISCOVERY January 2018 Conversion of PRPS Hexamer to Monomer by AMPK-Mediated Phosphorylation Inhibits Nucleotide

AMPK Phosphorylates PRPS to Reduce Nucleotide Synthesis RESEARCH ARTICLE

January 2018 CANCER DISCOVERY | 103

consumption of ATP for maintaining cell homeostasis. Animal studies showed that the expression of those non-phosphorylatable mutants inhibited brain tumorigenesis. Importantly, 2-DG treatment, which inhibits tumor cell gly-colysis, inhibited tumor growth, and that inhibition was enhanced by the expression of PRPS1 S180A and PRPS2 S183A. These findings highlight the significant role of reduced PRPS1/2 activity in counteracting energy stress.

The identified phosphorylation sequence of PRPS1 and PRPS2 is suboptimal to the AMPK recognition motif. How-ever, our in vitro and in vivo results supported the finding that AMPK directly phosphorylates PRPS1 S180 and PRPS2 S183. This finding, together with the reports that AMPK substrates such as histone H2B and eukaryotic elongation factor 2 kinase (eEF2K) contain a nonclassic AMPK recognition motif (36–38), indicate that AMPK tolerates some flexibility in its recognition motif for phosphorylation.

AMPK is a central metabolic switch that is found in all eukaryotes; it governs glucose and lipid metabolism in response to alterations in nutrients and intracellular energy

levels (39, 40). As a crucial cellular energy sensor, AMPK regulates energy balance at the cellular level by activating cata-bolic pathways, including glucose uptake, glycolysis, fatty-acid uptake and oxidation, and autophagy, and by inhibiting ana-bolic pathways, including the transcription of lipogenic and gluconeogenic enzymes and syntheses of fatty acid, triglycer-ide, cholesterol, glycogen, proteins, and ribosomal RNA (39, 40). Our findings elucidate yet another function of AMPK—direct regulation of nucleic acid synthesis in response to meta-bolic stress. Ultimately, therapeutic targeting of AMPK in the context of metabolic stress on tumor cells may be an effective strategy for disrupting PRPS1/2-mediated tumor growth.

METHODSMaterials

Primary rabbit polyclonal antibody against PRPS1/2 pS180/S183 was custom produced by Signalway Biotechnology. Mouse monoclo-nal antibodies against PRPS1/2 (sc-376440), GST (sc-138), α-tubulin (sc-5286), normal rabbit IgG (sc-2027), and normal mouse IgG

Figure 6.  AMPK-mediated PRPS1/2 phosphorylation copes with energy stress in tumorigenesis. A–D, 1 × 106 of parental U87 cells or U87 cells with knock-in of PRPS1/2-2S/A were intracranially injected into athymic nude mice (n = 7 per group). Two weeks after cell injection, 0.2 mL of 2-DG (500 mg/kg) was intraperitoneally administrated daily for 14 days. Mice were sacrificed and examined for tumor growth 28 days after injection. A, Representative hematoxylin and eosin–stained coronal brain sections are shown. Arrows point to tumor. B, Tumor volumes were calculated. Data represent the mean ± SD of seven mice. **, P < 0.001. C, Immunohistochemical analyses of xenograft brain tumor tissues were performed with an anti-PRPS1/2 pS180/S183 antibody. Scale bar, 100 μm. D, TUNEL analyses of the indicated tumor tissues were performed. Apoptotic cells were stained brown. Scale bar, 100 μm (left). Apoptotic cells were quantified, and the data represent the mean ± SD from 10 microscope fields (right). **, P < 0.001. E, Schematic model of AMPK-mediated nucleic acid and NAD synthesis under energy stress.

A

D E

B C

2-DG

2-DG

TU

NE

L-po

sitiv

e ce

lls (

%)

2-DG

90

2-DG−

2-DG−

**

PRPS

Nucleotide& NAD

synthesis

[PRPP]

Nucleotide& NAD

synthesis

[PRPP]

P P

PRPS

Energy-sufficient

AMPK AMPK

Energy-stress

Activehexamer

Inactivemonomer

**

2-DG+

2-DG+

+

****

8070605040302010

0

8070605040302010

0

Paren

tal

Paren

tal

Paren

tal

PRPS1/2

-2S/A

-C1

PRPS1/2

-2S/A

-C2

Paren

tal

PRPS1/2

-2S/A

-C1

PRPS1/2

-2S/A

-C2

Paren

tal

PRPS1/2

-2S/A

-C1

PRPS1/2

-2S/A

-C2

PRPS1/2

-2S/A

-C1

PRPS1/2

-2S/A

-C2

PRPS1/2

-2S/A

-C1

PRPS1/2

-2S/A

-C2

+

+

Research. on October 13, 2020. © 2018 American Association for Cancercancerdiscovery.aacrjournals.org Downloaded from

Published OnlineFirst October 26, 2017; DOI: 10.1158/2159-8290.CD-17-0712

Page 11: Conversion of PRPS Hexamer to Monomer by AMPK-Mediated … · 94 | CANCER DISCOVERY January 2018 Conversion of PRPS Hexamer to Monomer by AMPK-Mediated Phosphorylation Inhibits Nucleotide

Qian et al.RESEARCH ARTICLE

104 | CANCER DISCOVERY January 2018 www.aacrjournals.org

(sc-2025) were obtained from Santa Cruz Biotechnology. Rabbit mon-oclonal antibodies against AMPKα (#5831), AMPKα pT172 (#2535), ACC1 (#3676), and ACC1 pS79 (#11818) were purchased from Cell Signaling Technology. Mouse monoclonal anti-FLAG (F1804), 2-DG (D8375), gel filtration calibration markers kit (MWGF1000), R5P, ribulose-5-phosphate, NADH, phosphoenolpyruvate, pyruvate kinase, lactic dehydrogenase, myokinase, thiamine pyrophosphate, α-glycerophosphate dehydrogenase, triosephosphate isomerase, tran-sketolase, nicotinamide, thiazolyl blue, phenazine ethosulfate, glucose-6-phosphate dehydrogenase, glucose 6-phosphate, AMP, ADP, ATP, and NAC were purchased from Sigma. Anti-DYKDDDDK (FLAG) tag affinity gel was purchased from BioLegend. [γ-32P]ATP was purchased from MP Biochemicals. D-[6-14C]-glucose was purchased from Ameri-can Radiolabeled Chemicals. D-[U-13C6]-glucose, [13C2,15N]-glycine, and [13C5,15N4]-hypoxanthine were purchased from Cambridge Isotope Laboratories. D-ribulose-5-phosphate 3-epimerase was obtained from Abcam. Active AMPK proteins (#P47-10H) were obtained from Signal-Chem. A769662 was purchased from Tocris Bioscience. Compound C was obtained from Millipore.

Cell Culture and TransfectionU87, A549, BxPC-3, MDA-MB-231, Hep3B, Huh7, and 293T cells

were purchased from ATCC between 2003 and 2009. U251 cells were purchased from Sigma in 2010. NHA cells were obtained from Lonza in 2015. Cells were maintained in DMEM supplemented with 10% bovine calf serum (Hyclone). AMPKα1 and AMPKα2 double-knockout mouse embryonic fibroblasts (obtained from B. Viollet, Institut Cochin, Paris, France, in June 2015) were maintained in DMEM supplemented with 10% fetal bovine serum. HPDE cells (obtained from Ming-Sound Tsao, Princess Margaret Cancer Centre, University Health Network, Toronto, Canada, in August 2017) were maintained in Keratinocyte-SFM medium (Thermo Fisher, #17005042). U87 and U251 cell lines were authenticated using short tandem repeat profiling at The University of Texas MD Anderson Cancer Center (Houston, Texas) in November 2010. NHA, A549, BxPC-3, MDA-MB-231, Hep3B, and Huh7 cell lines were authenticated in September 2017. Mycoplasma was not detected in these cells.

Glucose deprivation was carried out by culturing cells in DMEM without glucose (Gibco, #11966), supplemented with dialyzed bovine calf serum or fetal bovine serum. Cells were plated at a density of 4 × 105 per 60-mm dish or 1 × 105 per well in a six-well plate 18 hours before transfection. The transfection procedure was performed as previously described (41).

DNA Construction and MutagenesisPolymerase chain reaction (PCR)–amplified human PRPS1 and

PRPS2 were cloned into pcDNA3.1/hygro(+)-FLAG or pColdI (His) vector, respectively. Human AMPKα1 was cloned into pcDNA3-HA vector. GST-AMPKα11–312 was obtained from Addgene (#27632). FLAG-PRPS1 S180A, S103T, and A190T, FLAG-PRPS2 S183A, His-PRPS1 S180A, His-PRPS2 S183A, and GST-AMPKα11–312 T172A were constructed using the QuikChange site-directed mutagenesis kit (Stratagene). shRNA-targeting PRPS1 (5′-GATAATATGATCTG CACCT-3′) was constructed into the pGIPZ vector.

Purification of Recombinant ProteinsWT His-PRPS1, His-PRPS1 S180A, WT His-PRPS2, and His-

PRPS2 S183A were expressed in bacteria and purified as described previously (41).

Immunoprecipitation and Immunoblot AnalysesThe extraction of proteins using a modified buffer (50 mmol/L

Tris-HCl [pH 7.5], 0.01% SDS, 1% Triton X-100, 150 mmol/L NaCl, 1 mmol/L dithiothreitol, 0.5 mmol/L EDTA, 100 μmol/L PMSF,

100 μmol/L leupeptin, 1 μmol/L aprotinin, 100 μmol/L sodium orthovanadate, 100 μmol/L sodium pyrophosphate, and 1 mmol/L sodium fluoride) from cultured cells was followed by immunopre-cipitation and immunoblot analyses using corresponding antibodies, as described previously (42).

Mass Spectrometry AnalysisImmunoprecipitated FLAG-PRPS1 protein from glucose-deprived

U87 cells was digested and analyzed by LC/MS-MS on an Orbitrap-Elite mass spectrometer (Thermo Fisher Scientific), as described previously (16).

Analysis of Intermediate Metabolites by LC/MS-MSTo determine the incorporation of glucose carbon (13C6-glucose)

into intracellular nucleotide metabolites, extracts were prepared and analyzed by high-resolution mass spectrometry (HRMS). Approxi-mately 80% confluent cells were seeded in 10 cm dishes in triplicate. For glucose labeling, cells were washed with glucose-free DMEM and incubated in fresh medium containing 13C6-glucose (10 mmol/L) for 30 minutes. For glycine or hypoxanthine labeling, cells were incubated with DMEM supplemented with [13C2,15N]-glycine (0.4 mmol/L) or [13C5,15N4]-hypoxanthine (0.1 mmol/L) for 2 hours. Nucleotide metabolites and NAD were extracted using 90/9/1 (v/v/v) acetonitrile/water/formic acid. Samples were centrifuged at 17,000 g for 5 minutes at 4°C, and supernatants were transferred to clean tubes, followed by evaporation to dryness using nitrogen. Samples were reconstituted in 0.2% ammonium hydroxide in ammonium acetate (10 mmol/L), then 10 μL was injected into a Thermo Scien-tific Vanquish liquid chromatography system containing a Thermo Hypercarb 100 × 3 mm 3 μm HPLC column heated to 35°C with mobile phase A (MPA) consisting of 0.2% ammonium hydroxide in ammonium acetate (10 mmol/L) and mobile phase B (MPB) consist-ing of 0.2% ammonium hydroxide in acetonitrile. Using a flow rate of 0.3 mL/minute, the gradient elution program was: 0 minute (0% MPB)–2.0 minutes (0% MPB)–15.0 minutes (30% MPB)–15.1 minutes (95% MPB)–20.0 minutes (95% MPB)–20.1 minutes (0% MPB)–25.0 minutes (STOP). Data were acquired using a Thermo Orbitrap Fusion Tribrid Mass Spectrometer via Selected Ion Mode (SIM) electrospray positive mode. For measurement of PRPP, metabolites were extracted by 75/25 (v/v) ethanol/10 mmol/L HEPES buffer, and followed by similar protocol for measuring nucleotides. Data acquisition was performed via SIM electrospray negative mode. Peak integration and area calculation were performed by using Thermo TraceFinder software.

To determine the abundance of intracellular PRPP, nucleotide intermediates, and NAD, approximately 80% confluent cells were seeded in 10 cm dishes in triplicate. Extracts were prepared and ana-lyzed by HRMS as described above.

In Vitro Kinase AssayThe kinase reactions were performed as described previously (43).

In brief, purified active AMPK proteins (including AMPKα1/β1/γ1) were incubated with purified WT His-PRPS1, His-PRPS1 S180A, WT His-PRPS2, or His-PRPS2 S183A (100 ng) in 25 μL of kinase buffer (50 mmol/L Tris-HCl [pH 7.5], 100 mmol/L KCl, 50 mmol/L MgCl2, 1 mmol/L Na3VO4, 1 mmol/L DTT, 5% glycerol, 0.2 mmol/L AMP, 0.5 mmol/L ATP, and 10 μCi [γ-32P]ATP) at 25°C for 1 hour. The reac-tion was terminated by adding SDS-PAGE loading buffer and heat-ing at 100°C for 5 minutes. The reaction mixture was then subjected to SDS-PAGE analyses.

For stoichiometry measurement of PRPS1 or PRPS2 phospho-rylation by AMPK, a modified method was performed, as described previously (44). Active AMPK protein (25 ng) was incubated with purified His-PRPS1 or His-PRPS2 (200 ng) in 25 μL of kinase buffer

Research. on October 13, 2020. © 2018 American Association for Cancercancerdiscovery.aacrjournals.org Downloaded from

Published OnlineFirst October 26, 2017; DOI: 10.1158/2159-8290.CD-17-0712

Page 12: Conversion of PRPS Hexamer to Monomer by AMPK-Mediated … · 94 | CANCER DISCOVERY January 2018 Conversion of PRPS Hexamer to Monomer by AMPK-Mediated Phosphorylation Inhibits Nucleotide

AMPK Phosphorylates PRPS to Reduce Nucleotide Synthesis RESEARCH ARTICLE

January 2018 CANCER DISCOVERY | 105

(50 mmol/L Tris-HCl [pH 7.5], 100 mmol/L KCl, 50 mmol/L MgCl2, 1 mmol/L Na3VO4, 1 mmol/L DTT, 5% glycerol, 0.2 mmol/L AMP). The reaction was started by adding [γ-32P]ATP (0.1 mmol/L, 1000 cpm/pmol) at 25°C. At various time points, samples were taken for SDS-PAGE analysis. Protein bands corresponding to PRPS1 or PRPS2 were cut directly from the gel and dissolved in vials containing 500 μL of 3% (w/w) H2O2 by heating for 2 hours at 80°C. 32P incor-poration was quantified by Cerenkov counting in a Beckman LS6500 scintillation counter and converted to moles of Pi incorporated per mole of PRPS1 or PRPS2.

CRISPR/Cas9-Mediated Genome EditingGenomic mutations were introduced into cells using the CRISPR/

Cas9 system, as described previously (45). Single-guide RNAs (sgRNA) were designed to target the genomic area adjacent to the PRPS1/2 mutation sites using the CRISPR design tool (http://crispr.mit.edu/). The annealed guide RNA oligonucleotides were inserted into PX458 vector (Addgene) digested with the BbsI restric-tion enzyme (46). Cells were seeded at 60% confluence, followed by cotransfection of sgRNAs (0.5 μg) and single-stranded donor oligonucleotide (10 pmol) as a template to introduce mutations. Twenty-four hours after transfection, cells were trypsinized and diluted for single cells and seeded into 96-well plates. Genomic DNA was extracted from GFP-positive cells, followed by sequencing of the PCR products spanning the mutation sites. The oligonucleotide and primers used for sgRNA cloning and genomic DNA sequencing are listed in Supplementary Table S1.

Size-Exclusion ChromatographySize-exclusion chromatography was performed using a HiPrep

16/60 Sephacryl S-300 HR column (GE), as described previously (47). Cell lysates (1 mL of 20 mg) or purified proteins (20 μg) were loaded onto the column and fractionated at a flow rate of 0.5 mL/minute. After passing through the void volume, the sample fractions were collected (1 mL for each) and analyzed by immunoblotting assay. The standard curve of the elution was plotted against LogMW by using a size-exclusion chromatography calibration marker kit (Sigma) according to the vendor’s instructions.

Influx of Glucose into Nucleic AcidsAn assay of the influx of glucose to nucleic acids was performed

as described previously, with modifications (16). In brief, cells were seeded in a six-well plate and incubated in 2 mL of fresh medium spiked with 1 μCi of D-[6-14C] glucose (0.01 mmol/L) for 0.5 hours, followed by washing with PBS. RNA or DNA was extracted, and 14C-RNA or 14C-DNA was assayed using a scintillation counter.

RT-PCR Analysis of PRPS1/2 ExpressionThe expression pattern of PRPS in cells was identified as described

previously (16). In brief, total RNA was extracted using TRIzol reagent (Invitrogen) according to the manufacturer’s instructions. A cDNA library was generated by reverse-transcribing 1 μg of RNA using an iScript cDNA synthesis kit (Bio-Rad) and amplified with a pair of primers (listed in Supplementary Table S1) that recognized both PRPS1 and PRPS2 cDNA. For distinguishing the expression of PRPS1/2, 1 μg of PCR products was digested by HindIII, and the products of digestion were resolved on a 1.5% agarose gel.

Measurement of PRPS1/2 ActivityPRPS1/2 activity was determined as described previously (16). In

brief, purified recombinant or immunoprecipitated WT or mutant PRPS1 or PRPS2 (10 ng) was incubated in 100 μL of reaction buffer (50 mmol/L Tris-HCl [pH 7.5], 4.7 mmol/L R5P, 0.4 mmol/L NADH, 3.2 mmol/L ATP, 1.8 mmol/L phosphoenolpyruvate, 6 mmol/L

MgCl2, 31 mmol/L NaHCO3, 7 U pyruvate kinase, 10 U lactic dehy-drogenase, and 10 U myokinase) at 25°C in a 96-well plate. The absorbance was read at 340 nm in kinetic mode for 5 minutes using a multidetection microplate reader (BMG Labtech).

Measurement of Intracellular ATP LevelsCells (1 × 106) were washed and lysed in ATP assay buffer. Samples

were deproteinized using 10 kDa spin columns (EMD Millipore). ATP levels were then determined using a BioVision ATP assay kit.

Measurement of Intracellular R5P LevelsThe intracellular concentration of R5P was measured using previ-

ously described methods, with some modifications (48). In brief, 1 × 106 cells were washed, collected, and homogenized in 1 mL of hypotonic lysis buffer. The homogenates were centrifuged at 4°C for 10 minutes at maximum speed, and the supernatants were deprotein-ized using 10 kDa spin columns (EMD Millipore). The flow-through was mixed with 2 × reaction buffer (100 mmol/L Tris-HCl [pH 8.0], 3.4 mmol/L ribulose-5-P, 0.004% w/v thiamine pyrophosphate, 30 mmol/L MgCl2, 0.4 mmol/L NADH, 1 U α-glycerophosphate dehydro-genase, 10 U triosephosphate isomerase, and 1 U transketolase). The reaction was initiated by adding D-ribulose-5-phosphate 3-epimerase (0.5 U) and incubated at 25°C for 30 minutes. The absorbance was read at 340 nm using a multidetection microplate reader.

Measurement of Intracellular NADPH and NADP+ LevelsThe intracellular levels of NADPH and total NADP (NADPH +

NADP+) were measured according to previously described methods (31, 49). In brief, 2 × 106 cells were lysed in 400 μL of extraction buffer containing 20 mmol/L nicotinamide, 20 mmol/L NaHCO3, and 100 mmol/L Na2CO3 and centrifuged for 10 minutes at maxi-mum speed. The supernatant was kept on ice in the dark until use. For extracting NADPH, 150 μL of the supernatant was incubated at 60°C for 30 minutes because heating destroys the oxidized form (NADP+) while having no effect on the reduced form (NADPH). Next, 160 μL of NADP-cycling buffer (100 mmol/L Tris-HCl [pH 8.0], 0.5 mmol/L thiazolyl blue [MTT], 2 mmol/L phenazine etho-sulfate, and 5 mmol/L EDTA) containing glucose-6-phosphate dehydrogenase (1.3 U) was added to a 96-well plate containing 20 μL of sample. After 1 minute incubation in the dark at 30°C, 20 μL of glucose 6-phosphate (10 mmol/L) was added to the mixture, and the change in absorbance at 570 nm was measured every 30 seconds for 4 minutes at 30°C with a microplate reader. The concentration of NADP+ was calculated by subtracting NADPH from the total NADP.

Measurement of Intracellular ROSThe fluorescence probe DCFDA (Abcam) was used to measure intra-

cellular ROS following the manufacturer’s instructions. In brief, 2 × 104 cells were seeded in a clear-bottom 96-well plate. After treatment, cells were washed with Hank’s Balanced Salt Solution (HBSS), followed by incubation with DCFDA (25 μmol/L)-containing HBSS for 30 minutes at 37°C in the dark. Cells were washed and green fluorescence was meas-ured with a Synergy HT microplate reader (BioTek) at 485/535 nm.

Apoptosis AnalysisCells were stained with the Annexin V-Alexa Fluor 488 apopto-

sis detection kit (ThermoFisher), followed by cytometry (BD Bio-sciences).

TUNEL AssayMouse tumor tissues were sectioned at 5 μm thickness. Apoptotic

cells were counted using the DeadEnd Colorimetric TUNEL System (Promega) according to the manufacturer’s instructions.

Research. on October 13, 2020. © 2018 American Association for Cancercancerdiscovery.aacrjournals.org Downloaded from

Published OnlineFirst October 26, 2017; DOI: 10.1158/2159-8290.CD-17-0712

Page 13: Conversion of PRPS Hexamer to Monomer by AMPK-Mediated … · 94 | CANCER DISCOVERY January 2018 Conversion of PRPS Hexamer to Monomer by AMPK-Mediated Phosphorylation Inhibits Nucleotide

Qian et al.RESEARCH ARTICLE

106 | CANCER DISCOVERY January 2018 www.aacrjournals.org

Structural View of PRPS1The structural view of PRPS1 hexamer (PDB: 2H06) was demon-

strated by Cn3D macromolecular structure viewer (https://www.ncbi.nlm.nih.gov/Structure/CN3D/cn3d.shtml; ref. 50), and the view of PRPS1 dimer was displayed by USFC Chimera (https://www.cgl.ucsf.edu/chimera/; ref. 51).

Cell Viability AnalysisCells (2 × 105) were plated in DMEM with 10% FBS. The viable

cells were stained with trypan blue (0.5%) and counted using Beck-man Coulter.

Intracranial InjectionCells (1 × 106) suspended in 5 μL of DMEM were intracranially

injected into 4-week-old female athymic nude mice, as described previously (52). Seven mice were included in each group. 2-DG treat-ment was initiated at the fourteenth day after the intracranial injec-tion of tumor cells. 2-DG (0.2 mL, 500 mg/kg) was intraperitoneally administered daily for 14 days. Mice were sacrificed 28 days after injection. Their brains were harvested, fixed in formaldehyde (4%), and embedded in paraffin. Tumor formation was determined by hematoxylin and eosin staining. Tumor volume was calculated by 0.5 × L × W2 (L, length; W, width). The use of animals was approved by the Institutional Review Board of MD Anderson Cancer Center.

Statistical AnalysisThe statistical analysis was conducted using the two-tailed

unpaired Student t test. All data represent the mean ± SD of three independent experiments/samples unless specifically indicated.

Disclosure of Potential Conflicts of InterestZ. Lu has ownership interest (including patents) in Signalway

Biotechnology. No potential conflicts of interest were disclosed by the other authors.

Authors’ ContributionsConception and design: X. Qian, Z. LuDevelopment of methodology: X. Qian, X. Li, J.-H. Lee, P.L. LorenziAcquisition of data (provided animals, acquired and managed patients, provided facilities, etc.): X. Qian, X. Li, L. Tan, Y. Xia, Q. Cai, Y. ZhengAnalysis and interpretation of data (e.g., statistical analysis, biostatistics, computational analysis): X. Qian, L. Tan, J.-H. Lee, P.L. LorenziWriting, review, and/or revision of the manuscript: L. Tan, Y. Xia, P.L. Lorenzi, Z. LuAdministrative, technical, or material support (i.e., reporting or organizing data, constructing databases): X. Qian, Y. XiaStudy supervision: X. Qian, H. Wang, Z. LuOther (proofreading): Y. Zheng

AcknowledgmentsWe thank Dr. Li Li (Clinical and Translational Proteomics Service

Center, Institute of Molecular Medicine, The University of Texas Health Center at Houston) for technical assistance, and Ann Sutton (The University of Texas MD Anderson Cancer Center) for her critical reading of the manuscript.

Grant SupportThis work was supported by National Institute of Neurological

Disorders and Stroke grant R01 NS089754 (Z. Lu), National Cancer Institute grants 2R01 CA109035 (Z. Lu) and R01 CA169603 (Z. Lu), 2P50 CA127001 (Brain Cancer SPORE), MD Anderson’s Cancer

Center Support Grant CA016672, and the National Natural Science Foundation of China (NSFC) 816228011 (H. Wang), 81572499, and 81572700. Z. Lu is a Ruby E. Rutherford Distinguished Professor.

The costs of publication of this article were defrayed in part by the payment of page charges. This article must therefore be hereby marked advertisement in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.

Received June 21, 2017; revised September 12, 2017; accepted October 19, 2017; published OnlineFirst October 26, 2017.

REFERENCES 1. Lyssiotis CA, Cantley LC. Acetate fuels the cancer engine. Cell

2014;159:1492–4. 2. Koppenol WH, Bounds PL, Dang CV. Otto Warburg’s contribu-

tions to current concepts of cancer metabolism. Nat Rev Cancer 2011;11:325–37.

3. Vander Heiden MG, Cantley LC, Thompson CB. Understanding the Warburg effect: the metabolic requirements of cell proliferation. Sci-ence 2009;324:1029–33.

4. Lyssiotis CA, Cantley LC. Metabolic syndrome: F stands for fructose and fat. Nature 2013;502:181–2.

5. Ishimoto T, Lanaspa MA, Le MT, Garcia GE, Diggle CP, Maclean PS, et  al. Opposing effects of fructokinase C and A isoforms on fructose-induced metabolic syndrome in mice. Proc Nat Acad Sci U S A 2012;109:4320–5.

6. Ben-Sahra I, Howell JJ, Asara JM, Manning BD. Stimulation of de novo pyrimidine synthesis by growth signaling through mTOR and S6K1. Science 2013;339:1323–8.

7. Ben-Sahra I, Hoxhaj G, Ricoult SJ, Asara JM, Manning BD. mTORC1 induces purine synthesis through control of the mitochondrial tet-rahydrofolate cycle. Science 2016;351:728–33.

8. Robitaille AM, Christen S, Shimobayashi M, Cornu M, Fava LL, Moes S, et  al. Quantitative phosphoproteomics reveal mTORC1 activates de novo pyrimidine synthesis. Science 2013;339:1320–3.

9. Kornberg A, Lieberman I, Simms ES. Enzymatic synthesis and prop-erties of 5-phosphoribosylpyrophosphate. J Biol Chem 1955;215: 389–402.

10. Khorana HG, Fernandes JF, Kornberg A. Pyrophosphorylation of ribose 5-phosphate in the enzymatic synthesis of 5-phosphorylribose 1-pyrophosphate. J Biol Chem 1958;230:941–8.

11. Hove-Jensen B. Mutation in the phosphoribosylpyrophosphate syn-thetase gene (prs) that results in simultaneous requirements for purine and pyrimidine nucleosides, nicotinamide nucleotide, histidine, and tryptophan in Escherichia coli. J Bacteriol 1988;170:1148–52.

12. Nosal JM, Switzer RL, Becker MA. Overexpression, purification, and characterization of recombinant human 5-phosphoribosyl-1-py-rophosphate synthetase isozymes I and II. J Biol Chem 1993;268: 10168–75.

13. Li S, Lu Y, Peng B, Ding J. Crystal structure of human phosphoribo-sylpyrophosphate synthetase 1 reveals a novel allosteric site. Biochem J 2007;401:39–47.

14. Chen P, Liu Z, Wang X, Peng J, Sun Q, Li J, et al. Crystal and EM struc-tures of human phosphoribosyl pyrophosphate synthase I (PRS1) provide novel insights into the disease-associated mutations. PloS One 2015;10:e0120304.

15. Li X, Qian X, Lu Z. Fructokinase A acts as a protein kinase to promote nucleotide synthesis. Cell Cycle 2016;15:2689–90.

16. Li X, Qian X, Peng LX, Jiang Y, Hawke DH, Zheng Y, et al. A splicing switch from ketohexokinase-C to ketohexokinase-A drives hepatocel-lular carcinoma formation. Nat Cell Biol 2016;18:561–71.

17. Li B, Li H, Bai Y, Kirschner-Schwabe R, Yang JJ, Chen Y, et al. Nega-tive feedback-defective PRPS1 mutants drive thiopurine resistance in relapsed childhood ALL. Nat Med 2015;21:563–71.

18. Cunningham JT, Moreno MV, Lodi A, Ronen SM, Ruggero D. Protein and nucleotide biosynthesis are coupled by a single rate-limiting enzyme, PRPS2, to drive cancer. Cell 2014;157:1088–103.

Research. on October 13, 2020. © 2018 American Association for Cancercancerdiscovery.aacrjournals.org Downloaded from

Published OnlineFirst October 26, 2017; DOI: 10.1158/2159-8290.CD-17-0712

Page 14: Conversion of PRPS Hexamer to Monomer by AMPK-Mediated … · 94 | CANCER DISCOVERY January 2018 Conversion of PRPS Hexamer to Monomer by AMPK-Mediated Phosphorylation Inhibits Nucleotide

AMPK Phosphorylates PRPS to Reduce Nucleotide Synthesis RESEARCH ARTICLE

January 2018 CANCER DISCOVERY | 107

19. Mannava S, Grachtchouk V, Wheeler LJ, Im M, Zhuang D, Slavina EG, et al. Direct role of nucleotide metabolism in C-MYC-dependent proliferation of melanoma cells. Cell Cycle 2008;7:2392–400.

20. McMahon SB. Control of nucleotide biosynthesis by the MYC onco-protein. Cell Cycle 2008;7:2275–6.

21. Lu Z, Hunter T. Degradation of activated protein kinases by ubiquit-ination. Ann Rev Biochem 2009;78:435–75.

22. Gwinn DM, Shackelford DB, Egan DF, Mihaylova MM, Mery A, Vasquez DS, et al. AMPK phosphorylation of raptor mediates a meta-bolic checkpoint. Mol Cell 2008;30:214–26.

23. Hardie DG, Schaffer BE, Brunet A. AMPK: an energy-sensing pathway with multiple inputs and outputs. Trends Cell Biol 2016;26:190–201.

24. Ingram-Smith C, Woods BI, Smith KS. Characterization of the acyl substrate binding pocket of acetyl-CoA synthetase. Biochemistry 2006;45:11482–90.

25. Crute BE, Seefeld K, Gamble J, Kemp BE, Witters LA. Functional domains of the alpha1 catalytic subunit of the AMP-activated protein kinase. J Biol Chem 1998;273:35347–54.

26. Canto C, Menzies KJ, Auwerx J. NAD(+) metabolism and the control of energy homeostasis: a balancing act between mitochondria and the nucleus. Cell Metab 2015;22:31–53.

27. Lane AN, Fan TW. Regulation of mammalian nucleotide metabolism and biosynthesis. Nucleic Acids Res 2015;43:2466–85.

28. Liu L, Ulbrich J, Muller J, Wustefeld T, Aeberhard L, Kress TR, et al. Deregulated MYC expression induces dependence upon AMPK-related kinase 5. Nature 2012;483:608–12.

29. Eguchi Y, Shimizu S, Tsujimoto Y. Intracellular ATP levels determine cell death fate by apoptosis or necrosis. Cancer Res 1997;57:1835–40.

30. Leist M, Single B, Castoldi AF, Kuhnle S, Nicotera P. Intracellular adenosine triphosphate (ATP) concentration: a switch in the decision between apoptosis and necrosis. J Exp Med 1997;185:1481–6.

31. Jeon SM, Chandel NS, Hay N. AMPK regulates NADPH homeosta-sis to promote tumour cell survival during energy stress. Nature 2012;485:661–5.

32. Yang M, Vousden KH. Serine and one-carbon metabolism in cancer. Nat Rev Cancer 2016;16:650–62.

33. Cairns RA, Harris IS, Mak TW. Regulation of cancer cell metabolism. Nat Rev Cancer 2011;11:85–95.

34. Tanaka H, Arakawa H, Yamaguchi T, Shiraishi K, Fukuda S, Matsui K, et al. A ribonucleotide reductase gene involved in a p53-dependent cell-cycle checkpoint for DNA damage. Nature 2000;404:42–9.

35. Jordan A, Reichard P. Ribonucleotide reductases. Ann Rev Biochem 1998;67:71–98.

36. Bungard D, Fuerth BJ, Zeng PY, Faubert B, Maas NL, Viollet B, et al. Signaling kinase AMPK activates stress-promoted transcription via histone H2B phosphorylation. Science 2010;329:1201–5.

37. Browne GJ, Finn SG, Proud CG. Stimulation of the AMP-activated protein kinase leads to activation of eukaryotic elongation factor 2

kinase and to its phosphorylation at a novel site, serine 398. J Biol Chem 2004;279:12220–31.

38. Johanns M, Pyr Dit Ruys S, Houddane A, Vertommen D, Herinckx G, Hue L, et al. Direct and indirect activation of eukaryotic elonga-tion factor 2 kinase by AMP-activated protein kinase. Cell Signall 2017;36:212–21.

39. Hardie DG, Ross FA, Hawley SA. AMPK: a nutrient and energy sensor that maintains energy homeostasis. Nat Rev Mol Cell Biol 2012;13:251–62.

40. Shackelford DB, Shaw RJ. The LKB1–AMPK pathway: metabo-lism and growth control in tumour suppression. Nat Rev Cancer 2009;9:563–75.

41. Xia Y, Wang J, Liu TJ, Yung WK, Hunter T, Lu Z. c-Jun downregu-lation by HDAC3-dependent transcriptional repression promotes osmotic stress-induced cell apoptosis. Mol Cell 2007;25:219–32.

42. Lu Z, Liu D, Hornia A, Devonish W, Pagano M, Foster DA. Activation of protein kinase C triggers its ubiquitination and degradation. Mol Cell Biol 1998;18:839–45.

43. Fang D, Hawke D, Zheng Y, Xia Y, Meisenhelder J, Nika H, et  al. Phosphorylation of beta-catenin by AKT promotes beta-catenin tran-scriptional activity. J Biol Chem 2007;282:11221–9.

44. Horman S, Morel N, Vertommen D, Hussain N, Neumann D, Beau-loye C, et  al. AMP-activated protein kinase phosphorylates and desensitizes smooth muscle myosin light chain kinase. J Biol Chem 2008;283:18505–12.

45. Li X, Jiang Y, Meisenhelder J, Yang W, Hawke DH, Zheng Y, et  al. Mitochondria-translocated PGK1 functions as a protein kinase to coordinate glycolysis and the TCA cycle in tumorigenesis. Mol Cell 2016;61:705–19.

46. Kleinstiver BP, Pattanayak V, Prew MS, Tsai SQ, Nguyen NT, Zheng Z, et  al. High-fidelity CRISPR-Cas9 nucleases with no detectable genome-wide off-target effects. Nature 2016;529:490–5.

47. Yang W, Lu Z. Nuclear PKM2 regulates the Warburg effect. Cell Cycle 2013;12:3154–8.

48. Lin R, Elf S, Shan C, Kang HB, Ji Q, Zhou L, et al. 6-Phosphogluconate dehydrogenase links oxidative PPP, lipogenesis and tumour growth by inhibiting LKB1–AMPK signalling. Nat Cell Biol 2015;17:1484–96.

49. Wagner TC, Scott MD. Single extraction method for the spectropho-tometric quantification of oxidized and reduced pyridine nucleotides in erythrocytes. Anal Biochem 1994;222:417–26.

50. Wang Y, Geer LY, Chappey C, Kans JA, Bryant SH. Cn3D: sequence and structure views for Entrez. Trends Biochem Sci 2000;25:300–2.

51. Pettersen EF, Goddard TD, Huang CC, Couch GS, Greenblatt DM, Meng EC, et al. UCSF Chimera–a visualization system for exploratory research and analysis. J Comput Chem 2004;25:1605–12.

52. Yang W, Xia Y, Ji H, Zheng Y, Liang J, Huang W, et al. Nuclear PKM2 regulates beta-catenin transactivation upon EGFR activation. Nature 2011;480:118–22.

Research. on October 13, 2020. © 2018 American Association for Cancercancerdiscovery.aacrjournals.org Downloaded from

Published OnlineFirst October 26, 2017; DOI: 10.1158/2159-8290.CD-17-0712

Page 15: Conversion of PRPS Hexamer to Monomer by AMPK-Mediated … · 94 | CANCER DISCOVERY January 2018 Conversion of PRPS Hexamer to Monomer by AMPK-Mediated Phosphorylation Inhibits Nucleotide

2018;8:94-107. Published OnlineFirst October 26, 2017.Cancer Discov   Xu Qian, Xinjian Li, Lin Tan, et al.   Energy StressPhosphorylation Inhibits Nucleotide Synthesis in Response to Conversion of PRPS Hexamer to Monomer by AMPK-Mediated

  Updated version

  10.1158/2159-8290.CD-17-0712doi:

Access the most recent version of this article at:

  Material

Supplementary

  http://cancerdiscovery.aacrjournals.org/content/suppl/2017/10/26/2159-8290.CD-17-0712.DC1

Access the most recent supplemental material at:

   

   

  Cited articles

  http://cancerdiscovery.aacrjournals.org/content/8/1/94.full#ref-list-1

This article cites 52 articles, 18 of which you can access for free at:

  Citing articles

  http://cancerdiscovery.aacrjournals.org/content/8/1/94.full#related-urls

This article has been cited by 2 HighWire-hosted articles. Access the articles at:

   

  E-mail alerts related to this article or journal.Sign up to receive free email-alerts

  SubscriptionsReprints and

  [email protected] at

To order reprints of this article or to subscribe to the journal, contact the AACR Publications

  Permissions

  Rightslink site. Click on "Request Permissions" which will take you to the Copyright Clearance Center's (CCC)

.http://cancerdiscovery.aacrjournals.org/content/8/1/94To request permission to re-use all or part of this article, use this link

Research. on October 13, 2020. © 2018 American Association for Cancercancerdiscovery.aacrjournals.org Downloaded from

Published OnlineFirst October 26, 2017; DOI: 10.1158/2159-8290.CD-17-0712