Advertisement

Getting all stressed out by vitamin C

Few experimental cancer therapies have incited as much debate as vitamin C. Yet the mechanistic effect of vitamin C on cancer cells is still poorly understood. Yun et al. studied human colorectal cancer cells with KRAS or BRAF mutations and found that they “handle” vitamin C in a different way than other cells, ultimately to their detriment (see the Perspective by Reczek and Chandel). Because a certain receptor is up-regulated in the mutant cells, they take up the oxidized form of vitamin C (dehydroascorbate). This leads to oxidative stress, inactivation of a glycolytic enzyme required by the mutant cells for growth, and finally cell death. Whether the selective toxicity of vitamin C to these mutant cells can be exploited therapeutically remains unclear.
Science, this issue p. 1391; see also p. 1317

Abstract

More than half of human colorectal cancers (CRCs) carry either KRAS or BRAF mutations and are often refractory to approved targeted therapies. We found that cultured human CRC cells harboring KRAS or BRAF mutations are selectively killed when exposed to high levels of vitamin C. This effect is due to increased uptake of the oxidized form of vitamin C, dehydroascorbate (DHA), via the GLUT1 glucose transporter. Increased DHA uptake causes oxidative stress as intracellular DHA is reduced to vitamin C, depleting glutathione. Thus, reactive oxygen species accumulate and inactivate glyceraldehyde 3-phosphate dehydrogenase (GAPDH). Inhibition of GAPDH in highly glycolytic KRAS or BRAF mutant cells leads to an energetic crisis and cell death not seen in KRAS and BRAF wild-type cells. High-dose vitamin C impairs tumor growth in Apc/KrasG12D mutant mice. These results provide a mechanistic rationale for exploring the therapeutic use of vitamin C for CRCs with KRAS or BRAF mutations.
Activating KRAS and BRAF mutations are found in approximately 40% and 10% of human colorectal cancers (CRCs), respectively (1). BRAF is a direct target of KRAS, and both activate the mitogen-activated protein kinase (MAPK) pathway. Clinical studies indicate that activating mutations in KRAS and BRAF predict resistance to epidermal growth factor receptor (EGFR)–targeting agents (24). Thus, novel therapies for KRAS or BRAF mutant CRCs are urgently needed.
Glucose uptake, as measured by [18F]fluorodeoxyglucose positron emission tomography (FDG-PET), correlates with KRAS or BRAF mutations and GLUT1 overexpression in CRCs (5, 6), consistent with our previous finding that KRAS or BRAF mutant CRC cells rewire glucose metabolism, in part by up-regulating GLUT1 expression (7). These data suggest a strategy for targeting KRAS or BRAF mutant cancers by exploiting the selective expression of GLUT1 and the metabolic liability that comes with increased reliance on glycolysis.
Dietary vitamin C is transported across cellular membranes by sodium vitamin C cotransporters (SVCTs) and facilitative glucose transporters (GLUTs) (8, 9). Whereas SVCTs transport vitamin C directly into the cell, GLUTs—mainly GLUT1 and GLUT3—transport the oxidized form of vitamin C, dehydroascorbate (DHA). After import, DHA is reduced to vitamin C at the expense of glutathione (GSH), thioredoxin, and nicotinamide adenine dinucleotide phosphate (NADPH) (10). Given that GLUT1 levels in KRAS and BRAF mutant cells are elevated, we hypothesized that the increase in DHA uptake could disrupt redox homeostasis and compromise cellular viability. To test our hypothesis, we used a panel of isogenic CRC cell lines harboring wild-type or mutant alleles of KRAS (HCT116 and DLD1) or BRAF (VACO432 and RKO) (7).
In cell culture media, vitamin C is oxidized to DHA (half-life ~70 min) unless reducing agents are added (fig. S1) (11). Using 14C-radiolabeled vitamin C, we tested which form of vitamin C (reduced or oxidized DHA) is preferentially imported. Both HCT116 and VACO432 cells took up [14C]vitamin C efficiently (Fig. 1A). However, adding GSH to the media to prevent oxidation of vitamin C to DHA abrogated [14C]vitamin C uptake (Fig. 1A). Furthermore, [14C]vitamin C uptake was significantly decreased in both HCT116 and VACO432 cells treated with a GLUT1-specific inhibitor, STF31, and in GLUT1 knockout cells (Fig. 1, A and B). Glucose competed with DHA for uptake in CRC cells (fig. S2). These results indicate that CRC cells preferentially import DHA, rather than vitamin C, and that this uptake is mediated by GLUT1 (fig. S3, A and B).
Fig. 1 KRAS and BRAF mutant cells predominantly take up DHA, the oxidized form of vitamin C, via GLUT1.
(A) DHA, but not vitamin C, is transported into CRC cells via GLUT1. [14C]Vitamin C was added to the culture media (2 mM glucose) for 30 min, followed by measurement of scintillation counts per minute (CPM) per microgram of protein input. Treating cells with GSH or STF31 (GLUT1 inhibitor) significantly reduced vitamin C uptake in all cases when compared to no GSH or STF31 treatment. One-way analysis of variance (ANOVA) followed by Dunnett’s posttest for multiple comparisons. *P < 0.01, **P < 0.001, n = 3. (B) [14C]Vitamin C uptake was monitored in 2 mM glucose and signal normalized to total protein. P, parental cells; WT-GLUT1, exogenously expressed GLUT1 in wild-type cells; GLUT1 KO, GLUT1 knockout cells. Asterisks indicate significant decreases in vitamin C uptake of wild-type or GLUT1 KO cells relative to the parental lines, KRAS or BRAF mutant cells (MUT), and WT-GLUT1. One-way ANOVA followed by Dunnett’s posttest. *P < 0.01, n = 3. (C) LC/MS analysis of intracellular vitamin C and DHA in KRAS or BRAF isogenic cell lines derived from HCT116 and VACO432, respectively. Cells were treated with 1 mM (HCT116) or 2 mM (VACO432) vitamin C for 1 hour before extracting vitamin C and DHA (Student’s t test, n = 6). All data represent means ± SD.
Given the increased expression of GLUT1 in mutant cells, we investigated whether KRAS or BRAF mutations influenced vitamin C uptake. We found that the mutant lines took up significantly more [14C]vitamin C than did their wild-type counterparts (Fig. 1, B and C). Overexpression of GLUT1 in wild-type cells was sufficient to increase [14C]vitamin C uptake to levels commensurate with those of the mutants (Fig. 1B and fig. S3C). Moreover, KRAS and BRAF mutant cells imported DHA faster than they did [14C]vitamin C (fig. S4), consistent with the observation that vitamin C must first be oxidized to DHA to enter cells through GLUT1. Together, these results indicate that GLUT1 is the primary means of vitamin C uptake in CRC cells and that elevated GLUT1 expression in KRAS or BRAF mutant cells drives increased DHA uptake.
We next asked whether the increased uptake of DHA in KRAS and BRAF mutant cells could affect their survival and growth. When plated at a low density and grown in low-glucose medium (2 mM), all cell lines grew at similar rates and formed colonies (fig. S5). However, 24 to 48 hours of vitamin C treatment inhibited KRAS and BRAF mutant cell growth and colony formation, with reduced effects on their wild-type counterparts (Fig. 2A and fig. S5). Because of the competitive nature of DHA import, mutant lines were most sensitive to vitamin C under low-glucose conditions (2 mM). Nonetheless, selective cytotoxicity against the mutant lines was achieved even under higher-glucose conditions (5 to 20 mM) when treating with less than 1 mM vitamin C (fig. S6), indicating that vitamin C can selectively kill mutant cells under physiological glucose concentration (5 to 10 mM).
Fig. 2 Vitamin C is selectively toxic to cells with mutant KRAS or BRAF alleles.
(A) Cell viability assay in 2 mM glucose or 2 mM glucose plus GSH in the presence of vitamin C (VC) for 48 hours (0.125 mM HCT116, DLD1, or RKO; 0.375 mM VACO432) after cells were plated at a low density. Values were normalized to vehicle control. Parental (P) and KRAS or BRAF mutant cells were significantly more sensitive than wild-type cells in the presence of vitamin C. One-way ANOVA with Dunnett’s posttest. *P < 0.0001, n = 3. (B) HCT116 (KRAS: G13D/+) or VACO432 (BRAF: V600E/+) cells were injected subcutaneously into the flank of 6- to 8-week-old female athymic nude mice (G13D, Gly-to-Asp mutations at codon 13; V600E, Val-to-Glu mutations at codon 600). After 7 to 10 days, mice were randomly divided into two groups. One group was treated with freshly prepared vitamin C in 400 μl of PBS (4 g/kg) twice a day via ip injection (HCT116, n = 6; VACO432, n = 6). Control group mice were treated with PBS with the same dosing schedule (HCT116, n = 4; VACO432, n = 7). Tumor sizes were measured two or three times per week in an unblinded manner. Experiments were repeated twice independently. (C) At 7 weeks of age, Apcflox/flox mice and Apcflox/flox/LSL-KrasG12D mice were treated with a single ip injection of low-dose tamoxifen (20 mg/kg) to activate the stem cell–specific Cre and facilitate loss of Apc and activation of the Kras G12D allele. Three weeks after tamoxifen injection, Apcflox/flox mice (8 male, 9 female) and Apcflox/flox/LSL-KrasG12D mice (7 male, 9 female) were divided into two groups [vitamin C (4 g/kg) or PBS] and treated daily with ip injections (five or six times per week). As a result of weight loss and the increased level of fecal occult blood as measured by the Hemoccult II SENSA test, all Apcflox/flox mice were killed at 6 weeks after treatment. Apcflox/flox/LSL-KrasG12D male mice were killed at 5 weeks after treatment and Apcflox/flox/LSL-KrasG12D female mice were killed at 7 weeks after treatment; average polyp numbers in the PBS group for female and male mice were similar. Apcflox/flox/LSL-KrasG12D mice experiments were repeated twice. Polyp numbers and volumes were determined in whole-mount tissue after methylene blue staining, using a dissecting microscope in an unblinded manner. (D) Immunoblots of GLUT1 protein, phospo-ERK1/2, and total ERK in tumors from Apcflox/flox mice (n = 4) and Apcflox/flox/LSL-KrasG12D mice (n = 4). Two separate polyps per mouse (pairs) were used for immunoblots. In. E., normal intestinal epithelial cells. (E) Absolute amounts of intracellular vitamin C were measured in tumors derived from Apcflox/flox mice and Apcflox/flox/LSL-KrasG12D mice treated with either vitamin C (4 g/kg) or PBS. Samples were harvested 1 hour after treatment. Two-way ANOVA (P = 0.0002) followed by Tukey’s test for multiple comparisons. All data represent means ± SD; n.s., not significant.
Plasma vitamin C concentrations greater than 10 mM are easily achieved in humans and in our murine pharmacokinetic study (fig. S7) without significant toxicity (12, 13). Vitamin C was cytotoxic rather than cytostatic, as evidenced by increased staining for the apoptotic marker annexin V in the mutants (fig. S8A). Adding GSH to the culture medium was sufficient to rescue the death of each mutant line (Fig. 2A). PIK3CA is one of three frequently mutated oncogenes in CRCs in addition to KRAS and BRAF. Unlike KRAS or BRAF, the PIK3CA genotype did not predict vitamin C sensitivity (fig. S8B). Notably, although overexpression of GLUT1 in wild-type cells increased vitamin C uptake (Fig. 1B), it did not sensitize wild-type cells to vitamin C (fig. S8C); these results indicate that high GLUT1 expression alone, without oncogene-induced metabolic reprograming, is not sufficient to make cells susceptible to vitamin C–dependent toxicity.
We next explored whether vitamin C altered the growth of KRAS and BRAF mutant CRC cells in mice. Mice bearing established xenografts derived from parental HCT116 and VACO432 cell lines were treated twice a day via intraperitoneal (ip) injection of high-dose vitamin C (4 g/kg) or phosphate-buffered saline (PBS; vehicle control) for 3 to 4 weeks, at which point control mice had to be killed because of tumor size. Vitamin C treatment significantly reduced tumor growth relative to vehicle control treatment (Fig. 2B). KRAS and BRAF wild-type isogenic HCT116 and VACO432 cell lines cannot form xenograft tumors in mice. To directly test the impact of Kras mutation on the sensitivity of tumors to vitamin C treatment, we generated a transgenic model of intestinal cancer, driven by either Apc mutation, or combined Apc and Kras (G12D) mutations. Compound mutant mice were generated by crossing available Apcflox mice (14), LSL-KrasG12D mice (15), and Lgr5-EGFP-creERT2 (16) animals, enabling intestinal restricted alteration of Apc and Kras. Tumors were induced with a single ip injection of low-dose tamoxifen (20 mg/kg) and treated daily thereafter with high-dose vitamin C (ip, 4 g/kg) for 5 to 7 weeks. Whereas Apcflox/flox mice showed no difference in polyp burden after vitamin C treatment, Apcflox/flox/KrasG12D mice had significantly fewer and smaller small intestine polyps (76 versus 165 in control group), confirming that vitamin C selectively affected Kras mutant tumors (Fig. 2C and fig. S9). Consistent with experiments in CRC lines, tumors from Apcflox/flox/KrasG12D mice showed higher GLUT1 expression and greater vitamin C uptake than did tumors from Apcflox/flox mice (Fig. 2, D and E, and fig. S10).
To investigate the mechanism by which vitamin C is selectively toxic to KRAS and BRAF mutant cells, we used liquid chromatography–tandem mass spectrometry (LC-MS/MS)–based metabolomics to profile metabolic changes after vitamin C treatment (17). In untreated KRAS and BRAF mutant lines, the relative intracellular metabolite levels of glycolysis and the nonoxidative arm of the pentose phosphate pathway (PPP) were increased relative to their isogenic wild-type counterparts (fig. S11). Addition of a MEK1/2 (MAPK kinase) inhibitor to the parental KRAS or BRAF mutant cells also decreased glycolytic and PPP metabolite levels, indicating that the increased metabolite levels were driven by oncogene-induced MAPK activity (fig. S12) (18). Notably, within 1 hour of vitamin C treatment, the metabolic profile of the mutant cells changed markedly. Glycolytic intermediates upstream of glyceraldehyde 3-phosphate dehydrogenase (GAPDH) accumulated while those downstream were depleted, which suggests that GAPDH was inhibited (Fig. 3A and fig. S13). Also, oxidative PPP metabolites increased (Fig. 3A and fig. S13), indicating that the blockage may shift glycolytic flux into the oxidative PPP. Indeed, vitamin C treatment stimulated oxidative PPP-dependent 14CO2 production from [1-14C]glucose in both KRAS and BRAF mutant cells, and to a lesser degree in wild-type cells (fig. S14A). Decreased NADPH/NADP+ ratios are known to activate glucose-6-phosphate dehydrogenase allosterically to enhance oxidative PPP flux. The increased flux is an attempt to restore cytosolic NADPH back to homeostasis to mitigate oxidative stress (19). We reasoned that DHA uptake may deplete cellular GSH and NADPH as they are consumed in reducing DHA to vitamin C. If the capacity of this pathway to restore GSH levels is exceeded, cellular reactive oxygen species (ROS) increase because GSH is the major cellular antioxidant (20). Indeed, the ratio of reduced to oxidized glutathione decreased as intracellular vitamin C increased (Fig. 3B and fig. S14B). Cysteine, the major limiting precursor for GSH biosynthesis, was also depleted after vitamin C treatment (fig. S13). As expected, vitamin C treatment induced a substantial increase in endogenous ROS in KRAS and BRAF mutant cells (Fig. 3C).
Fig. 3 Vitamin C inhibits glycolysis, thereby depleting ATP and selectively killing KRAS and BRAF mutant cells.
(A) Heat map depicting significantly changed glycolytic and pentose phosphate pathway (PPP) metabolite levels in mutant cells after 1 hour of vitamin C (VC) or vehicle (CON) treatment, as analyzed by LC-MS/MS. Red, increase; blue, decrease; TCA, tricarboxylic acid cycle. (B) Relative ratios of reduced to oxidized glutathione (GSH/GSSG) in KRAS and BRAF isogenic cell lines determined by LC-MS/MS as in (A). The ratio was significantly decreased after vitamin C treatment in KRAS or BRAF mutant cells as well as wild-type cells (Student’s t test, *P < 0.002, n = 3), but the extent was greater in the mutant cells than in the wild-type cells. (C) After 1 hour of vitamin C treatment, cells were incubated with the ROS-sensitive fluorescent dye DCF-DA for 30 min and fluorescence measured by flow cytometry. Asterisks indicate significant increases in ROS after vitamin C treatment (Student’s t test, *P < 0.01, n = 3). (D) The extracellular acidification rate (ECAR) was monitored in KRAS and BRAF isogenic cell lines. Red arrows indicate the time of vitamin C or vehicle (CON) addition (n = 6). (E) ATP levels were determined in KRAS and BRAF isogenic cell lines after 1 hour of vitamin C treatment. Although ATP levels were significantly decreased in all cells (Student’s t test, *P < 0.05, **P < 0.002, n = 3), the decrease was much more pronounced in KRAS or BRAF mutant cells (two-way ANOVA). (F) Cells were treated with vitamin C alone or combined with NAC for 1 hour before immunoblotting for Thr172 phosphorylation (p-AMPK) or total AMPK (t-AMPK). (G) Cells were treated with vitamin C alone or combined with NAC, pyruvate (Pyr), or Trolox for 48 hours and viability measured with a CellTiter-Glo assay (Promega). Cell viability in parental (P) and KRAS or BRAF mutant cells relative to wild-type cells was significantly decreased with vitamin C alone but not with vitamin C combination treatments. One-way ANOVA with Dunnett’s posttest. *P < 0.0001, n = 3. (H) Eight-week-old female athymic nude mice with subcutaneous tumors from parental HCT116 cells were treated with vitamin C alone (4 g/kg), NAC alone (30 mM in drinking water), vitamin C plus NAC, or PBS twice a day via ip injection. Tumor sizes were measured once per week in an unblinded manner. Experiments were repeated twice independently. Relative to PBS, vitamin C treatment alone significantly decreased tumor growth (P = 0.016), but adding NAC to the vitamin C treatment abolished this effect (P = 0.845). Mixed effect analysis followed by Tukey’s test. In (A) to (F), 1 and 2 mM vitamin C were used for HCT116 and VACO432 cells, respectively; for viability assays at low cell densities (G), 0.125 and 0.375 mM vitamin C were used for HCT116 and VACO432 cells, respectively. All data represent means ± SD.
Given that cancer cells with KRAS or BRAF mutations are heavily dependent on glycolysis for survival and growth, and that pyruvate (the end product of glycolysis) is a major carbon source for the mitochondrial TCA cycle (7, 21), we hypothesized that inhibition of GAPDH, a glycolytic enzyme,might deplete adenosine triphosphate (ATP) and thereby induce an energetic crisis ultimately leading to cell death. Vitamin C treatment caused a rapid decrease in the glycolytic rate in KRAS and BRAF mutant cells, but not in wild-type cells, as determined by the extracellular acidification rate, a proxy for lactate production (Fig. 3D and fig. S15). Accordingly, vitamin C induced a significant drop in ATP levels, with a concomitant increase in adenosine monophosphate (AMP) levels (Fig. 3E and fig. S16A). Within 1 hour, AMP-activated protein kinase (AMPK), a marker for energy stress, was activated; activation was strongest in the mutant lines (Fig. 3F). The cell-permeable reducing agent and glutathione precursor N-acetylcysteine (NAC) rescued both AMPK activation and cell death in the mutant lines (Fig. 3, F and G). Consistent with the in vitro results, supplementing drinking water with NAC over the course of vitamin C treatment abolished the ability of vitamin C to reduce xenograft growth (Fig. 3H). Similarly, pyruvate and oxaloacetate, both of which can enter the TCA cycle and thus provide ATP, rescued energy stress and cell death, as did Trolox (a water-soluble analog of the antioxidant vitamin E) (Fig. 3G and fig. S16, B and C). Rotenone, a complex I inhibitor, attenuated the ability of pyruvate to rescue vitamin C–induced cytotoxicity (fig. S17), indicating that the lack of mitochondrial substrates caused by glycolytic inhibition also contributes to ATP depletion in mutant cells (21).
We next sought to determine the mechanism by which vitamin C inhibits GAPDH. GAPDH is known to have an active-site cysteine (Cys152) that is targeted by ROS (22). The active-site cysteine can undergo reversible S-glutathionylation in which the oxidized cysteine forms a mixed disulfide with GSH (Cys-GSH), or undergo further irreversible oxidations that include sulfonic acid (Cys-SO3H) (23, 24). Both cases result in loss of GAPDH activity. We measured GAPDH S-glutathionylation after vitamin C treatment by immunoprecipitating endogenous GAPDH and blotting with an antibody that recognizes S-glutathionylation under nonreducing conditions. In both KRAS and BRAF mutant lines, GAPDH S-glutathionylation levels were higher in vitamin C–treated cells than in vehicle-treated cells by a factor of 2 to 3 (Fig. 4A). However, GAPDH sulfonylation was not detected with a GAPDH-SO3H antibody (Fig. 4B). GAPDH activity was assayed in lysates of vitamin C treated cells to confirm inhibition by S-glutathionylation (fig. S18). Treatment with vitamin C for 1 hour decreased GAPDH activity by 50% in both KRAS and BRAF mutant cells. Combining NAC with vitamin C fully rescued GAPDH activity (fig. S18).
Fig. 4 Vitamin C–induced ROS inhibits GAPDH by cysteine S-glutathionylation and depletion of NAD+.
(A) Cells were incubated with vehicle (CON) or vitamin C for 1 hour (HCT116, 1 mM; VACO432, 2 mM). Cell extracts were prepared in the presence of iodoacetic acid to prevent S-thiolation during extraction, immunoprecipitated with a GAPDH antibody, and analyzed by nonreducing SDS–polyacrylamide gel electrophoresis and probed with an antibody that recognizes S-glutathionylation. IP, immunoprecipitation; WB, Western blot. (B) HCT116 cells were incubated with vehicle, vitamin C, or H2O2 for 1 hour. Immunoprecipitations with a GAPDH antibody were performed the same way as in (A) and immunoblots were probed with a GAPDH-SO3H antibody. (C) Immunoblots for p(ADP)r (ADP-ribose polymers), Ser139-phosphorylated H2AX, total H2AX, and β-actin on lysates from cells treated with vehicle (CON) or vitamin C for 1 hour. (D) Cells were treated with vitamin C alone (0.125 mM) or vitamin C plus 10 μM olaparib (VC + PARPi) or 1 mM β-nicotinamide mononucleotide (VC + NMN). Viability after 48 hours of treatment was measured using a CellTiter-Glo assay and normalized to untreated controls. Asterisks indicate significant differences relative to KRAS or BRAF mutant cells treated with vitamin C alone. Two-way ANOVA followed by Tukey’s test. *P < 0.01, **P < 0.001, n = 3. (E) Schematic showing how vitamin C selectively kills KRAS or BRAF mutant cells.
We reasoned that the 50% reduction in GAPDH activity after vitamin C treatment could be explained by S-glutathionylation (Fig. 4A). However, given that the GAPDH substrates were added to the lysates to perform the activity assay, and in light of the striking accumulation (by as much as a factor of 19) of the GAPDH substrate glyceraldehyde-3-phosphate (G3P) (Fig. 3A and fig. S13), we suspected that additional mechanisms may contribute to GAPDH inhibition. This led us to examine the levels of the NAD+ substrate required for GAPDH-dependent oxidation of G3P. In contrast to G3P levels, intracellular NAD+ levels were significantly diminished after vitamin C treatment (fig. S19). PARP activation due to ROS-induced DNA damage consumes NAD+ to form adenosine diphosphate (ADP)–ribose polymers on acceptor proteins. We observed PARP activation and phosphorylation of H2AX, a marker of DNA damage, shortly after vitamin C treatment (Fig. 4C); this finding suggests that PARP activation may diminish NAD+ levels, thereby further inhibiting GAPDH activity by depleting substrate availability (25). To investigate whether PARP activation or NAD+ depletion contributes to vitamin C–induced cytotoxicity in KRAS and BRAF mutant cells, we treated cells with a PARP inhibitor, olaparib, or a cell-permeable NAD+ precursor, nicotinamide mononucleotide (NMN), before vitamin C treatment. Cell viability after vitamin C treatment was partially rescued by inhibiting PARP or supplementing with NMN (Fig. 4D). Taken together, these results indicate that in KRAS and BRAF mutant cells vitamin C–induced endogenous ROS inhibits GAPDH by both posttranslational modifications and NAD+ depletion, ultimately leading to an energetic crisis and cell death (Fig. 4E).
High-dose vitamin C cancer therapy has a controversial history. Although some early clinical studies indicated that vitamin C had antitumor activity (26, 27), others have shown little effect (28, 29). Recent studies suggest that the contradictory clinical data may be explained, at least in part, by differences in administration route; the millimolar vitamin C plasma concentrations cytotoxic to cancer cells are only achievable via intravenous administration, not via oral administration (30, 31). Given these findings, a growing number of phase I and phase II clinical trials are reevaluating intravenous infusion of vitamin C to treat various cancers (12, 13, 32, 33). However, despite the previous studies demonstrating that high-dose vitamin C is cytotoxic to cancer cells in vitro (3436) and that it delays tumor growth in xenograft models (37, 38), the mechanism by which vitamin C kills cancer cells while sparing normal cells has been unclear. Our findings address this fundamental question by suggesting that the oxidized form of vitamin C, DHA, is the pharmaceutically active agent, and that the selective toxicity of vitamin C to tumor cells stems from high GLUT1 expression combined with KRAS or BRAF oncogene-induced glycolytic addiction. Although it is unclear whether the results we have observed in our cell culture and mouse studies will translate to human tumors, our findings on the mechanism of action of vitamin C may warrant future investigation in clinical trials.

Acknowledgments

We thank B. Vogelstein and K. W. Kinzler for helpful suggestions; the Cantley lab members, B. Hopkins, F. Karreth, C. Lyssiotis, and G. DeNicola for comments on the manuscript; and M. Yuan, S. Breitkopf, J. Wong, and O. Mashadova for technical assistance. We apologize for publications not cited because of space limitations. L.C.C. owns equity in, receives compensation from, and serves on the board of directors and scientific advisory board of Agios Pharmaceuticals. Agios Pharmaceuticals is identifying metabolic pathways of cancer cells and developing drugs to inhibit such enzymes in order to disrupt tumor cell growth and survival. Supported by the Damon Runyon Cancer Research Foundation (J.Y. and I.I.C.C.), KL2 Career Development Awards (J.Y.), the U.S. Department of Defense (J.P.), National Cancer Institute grants P01 CA120964-07 and P01 CA117969-09 (L.C.C.), and NIH grant P01 CA120964 (J.M.A.). The authors declare no competing financial interests.

Supplementary Material

Summary

Materials and Methods
Figs. S1 to S19
References (39, 40)

Resources

File (yun.sm.pdf)

References and Notes

1
Sebolt-Leopold J. S., Herrera R., Targeting the mitogen-activated protein kinase cascade to treat cancer. Nat. Rev. Cancer 4, 937–947 (2004).
2
Lièvre A., Bachet J. B., Le Corre D., Boige V., Landi B., Emile J. F., Côté J. F., Tomasic G., Penna C., Ducreux M., Rougier P., Penault-Llorca F., Laurent-Puig P., KRAS mutation status is predictive of response to cetuximab therapy in colorectal cancer. Cancer Res. 66, 3992–3995 (2006).
3
Di Nicolantonio F., Martini M., Molinari F., Sartore-Bianchi A., Arena S., Saletti P., De Dosso S., Mazzucchelli L., Frattini M., Siena S., Bardelli A., Wild-type BRAF is required for response to panitumumab or cetuximab in metastatic colorectal cancer. J. Clin. Oncol. 26, 5705–5712 (2008).
4
Karapetis C. S., Khambata-Ford S., Jonker D. J., O’Callaghan C. J., Tu D., Tebbutt N. C., Simes R. J., Chalchal H., Shapiro J. D., Robitaille S., Price T. J., Shepherd L., Au H. J., Langer C., Moore M. J., Zalcberg J. R., K-ras mutations and benefit from cetuximab in advanced colorectal cancer. N. Engl. J. Med. 359, 1757–1765 (2008).
5
Kawada K., Nakamoto Y., Kawada M., Hida K., Matsumoto T., Murakami T., Hasegawa S., Togashi K., Sakai Y., Relationship between 18F-fluorodeoxyglucose accumulation and KRAS/BRAF mutations in colorectal cancer. Clin. Cancer Res. 18, 1696–1703 (2012).
6
Chen S. W., Chiang H. C., Chen W. T., Hsieh T. C., Yen K. Y., Chiang S. F., Kao C. H., Correlation between PET/CT parameters and KRAS expression in colorectal cancer. Clin. Nucl. Med. 39, 685–689 (2014).
7
Yun J., Rago C., Cheong I., Pagliarini R., Angenendt P., Rajagopalan H., Schmidt K., Willson J. K., Markowitz S., Zhou S., Diaz L. A., Velculescu V. E., Lengauer C., Kinzler K. W., Vogelstein B., Papadopoulos N., Glucose deprivation contributes to the development of KRAS pathway mutations in tumor cells. Science 325, 1555–1559 (2009).
8
Tsukaguchi H., Tokui T., Mackenzie B., Berger U. V., Chen X. Z., Wang Y., Brubaker R. F., Hediger M. A., A family of mammalian Na+-dependent L-ascorbic acid transporters. Nature 399, 70–75 (1999).
9
Vera J. C., Rivas C. I., Fischbarg J., Golde D. W., Mammalian facilitative hexose transporters mediate the transport of dehydroascorbic acid. Nature 364, 79–82 (1993).
10
Linster C. L., Van Schaftingen E., Vitamin C. Biosynthesis, recycling and degradation in mammals. FEBS J. 274, 1–22 (2007).
11
Vera J. C., Rivas C. I., Zhang R. H., Farber C. M., Golde D. W., Human HL-60 myeloid leukemia cells transport dehydroascorbic acid via the glucose transporters and accumulate reduced ascorbic acid. Blood 84, 1628–1634 (1994).
12
Stephenson C. M., Levin R. D., Spector T., Lis C. G., Phase I clinical trial to evaluate the safety, tolerability, and pharmacokinetics of high-dose intravenous ascorbic acid in patients with advanced cancer. Cancer Chemother. Pharmacol. 72, 139–146 (2013).
13
Hoffer L. J., Levine M., Assouline S., Melnychuk D., Padayatty S. J., Rosadiuk K., Rousseau C., Robitaille L., Miller W. H., Phase I clinical trial of i.v. ascorbic acid in advanced malignancy. Ann. Oncol. 19, 1969–1974 (2008).
14
Kuraguchi M., Wang X. P., Bronson R. T., Rothenberg R., Ohene-Baah N. Y., Lund J. J., Kucherlapati M., Maas R. L., Kucherlapati R., Adenomatous polyposis coli (APC) is required for normal development of skin and thymus. PLOS Genet. 2, e146 (2006).
15
Jackson E. L., Willis N., Mercer K., Bronson R. T., Crowley D., Montoya R., Jacks T., Tuveson D. A., Analysis of lung tumor initiation and progression using conditional expression of oncogenic K-ras. Genes Dev. 15, 3243–3248 (2001).
16
Barker N., Ridgway R. A., van Es J. H., van de Wetering M., Begthel H., van den Born M., Danenberg E., Clarke A. R., Sansom O. J., Clevers H., Crypt stem cells as the cells-of-origin of intestinal cancer. Nature 457, 608–611 (2009).
17
Yuan M., Breitkopf S. B., Yang X., Asara J. M., A positive/negative ion-switching, targeted mass spectrometry-based metabolomics platform for bodily fluids, cells, and fresh and fixed tissue. Nat. Protoc. 7, 872–881 (2012).
18
Ying H., Kimmelman A. C., Lyssiotis C. A., Hua S., Chu G. C., Fletcher-Sananikone E., Locasale J. W., Son J., Zhang H., Coloff J. L., Yan H., Wang W., Chen S., Viale A., Zheng H., Paik J. H., Lim C., Guimaraes A. R., Martin E. S., Chang J., Hezel A. F., Perry S. R., Hu J., Gan B., Xiao Y., Asara J. M., Weissleder R., Wang Y. A., Chin L., Cantley L. C., DePinho R. A., Oncogenic Kras maintains pancreatic tumors through regulation of anabolic glucose metabolism. Cell 149, 656–670 (2012).
19
Ralser M., Wamelink M. M., Kowald A., Gerisch B., Heeren G., Struys E. A., Klipp E., Jakobs C., Breitenbach M., Lehrach H., Krobitsch S., Dynamic rerouting of the carbohydrate flux is key to counteracting oxidative stress. J. Biol. 6, 10 (2007).
20
Pastore A., Federici G., Bertini E., Piemonte F., Analysis of glutathione: Implication in redox and detoxification. Clin. Chim. Acta 333, 19–39 (2003).
21
Weinberg S. E., Chandel N. S., Targeting mitochondria metabolism for cancer therapy. Nat. Chem. Biol. 11, 9–15 (2015).
22
Hwang N. R., Yim S. H., Kim Y. M., Jeong J., Song E. J., Lee Y., Lee J. H., Choi S., Lee K. J., Oxidative modifications of glyceraldehyde-3-phosphate dehydrogenase play a key role in its multiple cellular functions. Biochem. J. 423, 253–264 (2009).
23
Shenton D., Grant C. M., Protein S-thiolation targets glycolysis and protein synthesis in response to oxidative stress in the yeast Saccharomyces cerevisiae. Biochem. J. 374, 513–519 (2003).
24
Ravichandran V., Seres T., Moriguchi T., Thomas J. A., Johnston R. B., S-thiolation of glyceraldehyde-3-phosphate dehydrogenase induced by the phagocytosis-associated respiratory burst in blood monocytes. J. Biol. Chem. 269, 25010–25015 (1994).
25
Alano C. C., Garnier P., Ying W., Higashi Y., Kauppinen T. M., Swanson R. A., NAD+ depletion is necessary and sufficient for poly(ADP-ribose) polymerase-1-mediated neuronal death. J. Neurosci. 30, 2967–2978 (2010).
26
Cameron E., Pauling L., Supplemental ascorbate in the supportive treatment of cancer: Prolongation of survival times in terminal human cancer. Proc. Natl. Acad. Sci. U.S.A. 73, 3685–3689 (1976).
27
Cameron E., Pauling L., Supplemental ascorbate in the supportive treatment of cancer: Reevaluation of prolongation of survival times in terminal human cancer. Proc. Natl. Acad. Sci. U.S.A. 75, 4538–4542 (1978).
28
Creagan E. T., Moertel C. G., O’Fallon J. R., Schutt A. J., O’Connell M. J., Rubin J., Frytak S., Failure of high-dose vitamin C (ascorbic acid) therapy to benefit patients with advanced cancer. A controlled trial. N. Engl. J. Med. 301, 687–690 (1979).
29
Moertel C. G., Fleming T. R., Creagan E. T., Rubin J., O’Connell M. J., Ames M. M., High-dose vitamin C versus placebo in the treatment of patients with advanced cancer who have had no prior chemotherapy. A randomized double-blind comparison. N. Engl. J. Med. 312, 137–141 (1985).
30
Padayatty S. J., Sun H., Wang Y., Riordan H. D., Hewitt S. M., Katz A., Wesley R. A., Levine M., Vitamin C pharmacokinetics: Implications for oral and intravenous use. Ann. Intern. Med. 140, 533–537 (2004).
31
Levine M., Conry-Cantilena C., Wang Y., Welch R. W., Washko P. W., Dhariwal K. R., Park J. B., Lazarev A., Graumlich J. F., King J., Cantilena L. R., Vitamin C pharmacokinetics in healthy volunteers: Evidence for a recommended dietary allowance. Proc. Natl. Acad. Sci. U.S.A. 93, 3704–3709 (1996).
32
Monti D. A., Mitchell E., Bazzan A. J., Littman S., Zabrecky G., Yeo C. J., Pillai M. V., Newberg A. B., Deshmukh S., Levine M., Phase I evaluation of intravenous ascorbic acid in combination with gemcitabine and erlotinib in patients with metastatic pancreatic cancer. PLOS ONE 7, e29794 (2012).
33
Welsh J. L., Wagner B. A., van’t Erve T. J., Zehr P. S., Berg D. J., Halfdanarson T. R., Yee N. S., Bodeker K. L., Du J., Roberts L. J., Drisko J., Levine M., Buettner G. R., Cullen J. J., Pharmacological ascorbate with gemcitabine for the control of metastatic and node-positive pancreatic cancer (PACMAN): Results from a phase I clinical trial. Cancer Chemother. Pharmacol. 71, 765–775 (2013).
34
Chen Q., Espey M. G., Krishna M. C., Mitchell J. B., Corpe C. P., Buettner G. R., Shacter E., Levine M., Pharmacologic ascorbic acid concentrations selectively kill cancer cells: Action as a pro-drug to deliver hydrogen peroxide to tissues. Proc. Natl. Acad. Sci. U.S.A. 102, 13604–13609 (2005).
35
Tian W., Wang Y., Xu Y., Guo X., Wang B., Sun L., Liu L., Cui F., Zhuang Q., Bao X., Schley G., Chung T. L., Laslett A. L., Willam C., Qin B., Maxwell P. H., Esteban M. A., The hypoxia-inducible factor renders cancer cells more sensitive to vitamin C-induced toxicity. J. Biol. Chem. 289, 3339–3351 (2014).
36
Vuyyuri S. B., Rinkinen J., Worden E., Shim H., Lee S., Davis K. R., Ascorbic acid and a cytostatic inhibitor of glycolysis synergistically induce apoptosis in non-small cell lung cancer cells. PLOS ONE 8, e67081 (2013).
37
Kim J., Lee S. D., Chang B., Jin D. H., Jung S. I., Park M. Y., Han Y., Yang Y., Il Kim K., Lim J. S., Kang Y. S., Lee M. S., Enhanced antitumor activity of vitamin C via p53 in cancer cells. Free Radic. Biol. Med. 53, 1607–1615 (2012).
38
Ma Y., Chapman J., Levine M., Polireddy K., Drisko J., Chen Q., High-dose parenteral ascorbate enhanced chemosensitivity of ovarian cancer and reduced toxicity of chemotherapy. Sci. Transl. Med. 6, 222ra218 (2014).
39
Schepers A. G., Snippert H. J., Stange D. E., van den Born M., van Es J. H., van de Wetering M., Clevers H., Lineage tracing reveals Lgr5+ stem cell activity in mouse intestinal adenomas. Science 337, 730–735 (2012).
40
Hansler A., Chen Q., Gray J. D., Ross M. E., Finnell R. H., Gross S. S., Untargeted metabolite profiling of murine embryos to reveal metabolic perturbations associated with neural tube closure defects. Birth Defects Res. A 100, 623–632 (2014).

(0)eLetters

eLetters is a forum for ongoing peer review. eLetters are not edited, proofread, or indexed, but they are screened. eLetters should provide substantive and scholarly commentary on the article. Embedded figures cannot be submitted, and we discourage the use of figures within eLetters in general. If a figure is essential, please include a link to the figure within the text of the eLetter. Please read our Terms of Service before submitting an eLetter.

Log In to Submit a Response

No eLetters have been published for this article yet.

Information & Authors

Information

Published In

Science
Volume 350 | Issue 6266
11 December 2015

Article versions

You are viewing the most recent version of this article.

Submission history

Received: 15 December 2014
Accepted: 16 October 2015
Published in print: 11 December 2015

Permissions

Request permissions for this article.

Acknowledgments

We thank B. Vogelstein and K. W. Kinzler for helpful suggestions; the Cantley lab members, B. Hopkins, F. Karreth, C. Lyssiotis, and G. DeNicola for comments on the manuscript; and M. Yuan, S. Breitkopf, J. Wong, and O. Mashadova for technical assistance. We apologize for publications not cited because of space limitations. L.C.C. owns equity in, receives compensation from, and serves on the board of directors and scientific advisory board of Agios Pharmaceuticals. Agios Pharmaceuticals is identifying metabolic pathways of cancer cells and developing drugs to inhibit such enzymes in order to disrupt tumor cell growth and survival. Supported by the Damon Runyon Cancer Research Foundation (J.Y. and I.I.C.C.), KL2 Career Development Awards (J.Y.), the U.S. Department of Defense (J.P.), National Cancer Institute grants P01 CA120964-07 and P01 CA117969-09 (L.C.C.), and NIH grant P01 CA120964 (J.M.A.). The authors declare no competing financial interests.

Authors

Affiliations

Jihye Yun
Meyer Cancer Center, Department of Medicine, Weill Cornell Medical College, New York, NY 10065, USA.
Edouard Mullarky
Meyer Cancer Center, Department of Medicine, Weill Cornell Medical College, New York, NY 10065, USA.
Biological and Biomedical Sciences Graduate Program, Harvard Medical School, Boston, MA 02115, USA.
Changyuan Lu
Department of Pharmacology, Weill Cornell Medical College, New York, NY 10065, USA.
Kaitlyn N. Bosch
Meyer Cancer Center, Department of Medicine, Weill Cornell Medical College, New York, NY 10065, USA.
Adam Kavalier
Department of Pharmacology, Weill Cornell Medical College, New York, NY 10065, USA.
Keith Rivera
Cold Spring Harbor Laboratory, Cold Spring Harbor, NY 11724, USA.
Jatin Roper
Molecular Oncology Research Institute and Division of Gastroenterology, Tufts Medical Center, Boston, MA 02111, USA.
Iok In Christine Chio
Cold Spring Harbor Laboratory, Cold Spring Harbor, NY 11724, USA.
Eugenia G. Giannopoulou
Department of Physiology and Biophysics, Weill Cornell Medical College, New York, NY 10065, USA.
Present address: Department of Biological Sciences, New York City College of Technology, City University of New York, Brooklyn, NY 11201, USA, and Arthritis and Tissue Degeneration Program and David Z. Rosensweig Genomics Research Center, Hospital for Special Surgery, New York, NY 10021, USA.
Carlo Rago
Ludwig Center for Cancer Genetics and Therapeutics and Howard Hughes Medical Institute, Johns Hopkins Kimmel Cancer Center, Baltimore, MD 21231, USA.
Ashlesha Muley
Meyer Cancer Center, Department of Medicine, Weill Cornell Medical College, New York, NY 10065, USA.
John M. Asara
Division of Signal Transduction, Beth Israel Deaconess Medical Center and Department of Medicine, Harvard Medical School, Boston, MA 02115, USA.
Jihye Paik
Department of Pathology and Laboratory Medicine, Weill Cornell Medical College, New York, NY 10065, USA.
Olivier Elemento
Department of Physiology and Biophysics, Weill Cornell Medical College, New York, NY 10065, USA.
Zhengming Chen
Department of Biostatistics and Epidemiology, Weill Cornell Medical College, New York, NY 10065, USA.
Darryl J. Pappin
Cold Spring Harbor Laboratory, Cold Spring Harbor, NY 11724, USA.
Lukas E. Dow
Meyer Cancer Center, Department of Medicine, Weill Cornell Medical College, New York, NY 10065, USA.
Nickolas Papadopoulos
Ludwig Center for Cancer Genetics and Therapeutics and Howard Hughes Medical Institute, Johns Hopkins Kimmel Cancer Center, Baltimore, MD 21231, USA.
Steven S. Gross
Department of Pharmacology, Weill Cornell Medical College, New York, NY 10065, USA.
Lewis C. Cantley [email protected]
Meyer Cancer Center, Department of Medicine, Weill Cornell Medical College, New York, NY 10065, USA.

Notes

Corresponding author. E-mail: [email protected]

Metrics & Citations

Metrics

Article Usage

Altmetrics

Citations

Cite as

Export citation

Select the format you want to export the citation of this publication.

Cited by

  1. Resistance to targeted therapy in metastatic colorectal cancer: Current status and new developments, World Journal of Gastroenterology, 29, 6, (926-948), (2023).https://doi.org/10.3748/wjg.v29.i6.926
    Crossref
  2. Pleiotropic Functions of Nitric Oxide Produced by Ascorbate for the Prevention and Mitigation of COVID-19: A Revaluation of Pauling’s Vitamin C Therapy, Microorganisms, 11, 2, (397), (2023).https://doi.org/10.3390/microorganisms11020397
    Crossref
  3. High-Dose Ascorbate in Combination with Anti-PD1 Checkpoint Inhibition as Treatment Option for Malignant Melanoma, Cells, 12, 2, (254), (2023).https://doi.org/10.3390/cells12020254
    Crossref
  4. Overcoming EGFR Resistance in Metastatic Colorectal Cancer Using Vitamin C: A Review, Biomedicines, 11, 3, (678), (2023).https://doi.org/10.3390/biomedicines11030678
    Crossref
  5. Can plasma vitamin C predict survival in stage IV colorectal cancer patients? Results of a prospective cohort study, Frontiers in Nutrition, 10, (2023).https://doi.org/10.3389/fnut.2023.1110405
    Crossref
  6. Pharmacological vitamin C inhibits mTOR signaling and tumor growth by degrading Rictor and inducing HMOX1 expression, PLOS Genetics, 19, 2, (e1010629), (2023).https://doi.org/10.1371/journal.pgen.1010629
    Crossref
  7. Metabolic reprogramming in colorectal cancer: regulatory networks and therapy, Cell & Bioscience, 13, 1, (2023).https://doi.org/10.1186/s13578-023-00977-w
    Crossref
  8. Metabolism and Colorectal Cancer, Annual Review of Pathology: Mechanisms of Disease, 18, 1, (467-492), (2023).https://doi.org/10.1146/annurev-pathmechdis-031521-041113
    Crossref
  9. Understanding the rod-to-tube transformation of self-assembled ascorbyl dipalmitate lipid nanoparticles stabilized with PEGylated lipids, Nanoscale, 15, 6, (2602-2613), (2023).https://doi.org/10.1039/D2NR04987B
    Crossref
  10. Synergistic effect of glutathione and IgG4 in immune evasion and the implication for cancer immunotherapy, Redox Biology, 60, (102608), (2023).https://doi.org/10.1016/j.redox.2023.102608
    Crossref
  11. See more
Loading...

View Options

View options

PDF format

Download this article as a PDF file

Download PDF

Check Access

Log in to view the full text

AAAS ID LOGIN

AAAS login provides access to Science for AAAS Members, and access to other journals in the Science family to users who have purchased individual subscriptions.

Log in via OpenAthens.
Log in via Shibboleth.

More options

Register for free to read this article

As a service to the community, this article is available for free. Login or register for free to read this article.

Purchase this issue in print

Buy a single issue of Science for just $15 USD.

Media

Figures

Multimedia

Tables

Share

Share

Share article link

Share on social media