throbber
Vol 462 | 10 December 2009 | doi:10.1038/nature08617
`
`ARTICLES
`
`Cancer-associated IDH1 mutations
`produce 2-hydroxyglutarate
`
`Lenny Dang1, David W. White1, Stefan Gross1, Bryson D. Bennett2, Mark A. Bittinger1, Edward M. Driggers1,
`Valeria R. Fantin1, Hyun Gyung Jang1, Shengfang Jin1, Marie C. Keenan1, Kevin M. Marks1, Robert M. Prins3,
`Patrick S. Ward4, Katharine E. Yen1, Linda M. Liau3, Joshua D. Rabinowitz2, Lewis C. Cantley5, Craig B. Thompson4,
`Matthew G. Vander Heiden1{ & Shinsan M. Su1
`
`Mutations in the enzyme cytosolic isocitrate dehydrogenase 1 (IDH1) are a common feature of a major subset of primary
`human brain cancers. These mutations occur at a single amino acid residue of the IDH1 active site, resulting in loss of the
`enzyme’s ability to catalyse conversion of isocitrate to a-ketoglutarate. However, only a single copy of the gene is mutated in
`tumours, raising the possibility that the mutations do not result in a simple loss of function. Here we show that
`cancer-associated IDH1 mutations result in a new ability of the enzyme to catalyse the NADPH-dependent reduction of
`a-ketoglutarate to R(2)-2-hydroxyglutarate (2HG). Structural studies demonstrate that when arginine 132 is mutated to
`histidine, residues in the active site are shifted to produce structural changes consistent with reduced oxidative
`decarboxylation of isocitrate and acquisition of the ability to convert a-ketoglutarate to 2HG. Excess accumulation of 2HG
`has been shown to lead to an elevated risk of malignant brain tumours in patients with inborn errors of 2HG metabolism.
`Similarly, in human malignant gliomas harbouring IDH1 mutations, we find markedly elevated levels of 2HG. These data
`demonstrate that the IDH1 mutations result in production of the onco-metabolite 2HG, and indicate that the excess 2HG
`which accumulates in vivo contributes to the formation and malignant progression of gliomas.
`
`Mutations in the enzyme cytosolic isocitrate dehydrogenase 1 (IDH1)
`are found in approximately 80% of grade II–III gliomas and secondary
`glioblastomas in humans1–3. These mutations occur at a single amino
`acid residue of IDH1, arginine 132, which is most commonly mutated
`to histidine (R132H)1,3,4. Only a single copy of the gene has been found
`to be mutated in tumours1–6. Many of the high-grade gliomas with
`IDH1 mutations are secondary glioblastomas that have progressed
`from lower grade lesions1–3,5. When analysed in relation to other genes
`implicated in brain tumours, the compiled evidence suggests that
`IDH1 is often the first mutation that occurs2. Although these findings
`suggest that IDH1 mutations are selected for early during tumorigen-
`esis, why mutations in a single allele of IDH1 result in predilection for
`malignant progression is uncertain. It has been reported that the
`R132H mutation disrupts the ability of IDH1 to convert isocitrate
`to a-ketoglutarate3,7, but the consequences of this impaired enzymatic
`activity on cellular metabolism have not been systematically analysed.
`For example, although R132 IDH1 mutations might reduce the rate of
`cytosolic a-ketoglutarate production as suggested by others7, whether
`IDH1 mutations can influence the enzyme’s ability to act on a-
`ketoglutarate as a substrate has not been explored. This latter activity
`may be particularly important for the tumorigenic role of IDH1 muta-
`tions because cytosolic a-ketoglutarate is in equilibrium via transami-
`nation with glutamate that has a unique role in glia cell physiology,
`and IDH1 mutations are especially prevalent in malignant gliomas.
`
`IDH1 mutant expressing cells have elevated 2HG levels
`To understand the impact of IDH1 mutation on cellular metabolism,
`we profiled metabolites to identify changes in metabolite levels in
`
`cells expressing R132 mutant IDH1 compared with cells expressing
`wild-type IDH1. To initiate these studies, we stably transfected
`U87MG glioblastoma cells, which are wild-type for IDH1, with
`Myc-tagged wild-type or R132H mutant IDH1. Cells expressing either
`Myc-tagged wild-type or mutant IDH1 were used for metabolite pro-
`filing experiments (Fig. 1a). Metabolites extracted from exponentially
`growing cells were profiled by liquid chromatography–electrospray
`ionization–mass spectrometry (LC–MS). In initial survey analyses,
`full-scan LC–MS in negative ion mode (exact mass) was used to
`examine differences in metabolite species with an m/z between 110
`and 1,000. Relative quantitative data were collected for approximately
`850 ions and identities were proposed by comparison with known
`human metabolites. Identities of .100 species identified by a com-
`bination of exact mass and retention time match to purified standards
`were assigned8. There were no significant differences between cells
`expressing wild-type IDH1 when compared with parental cells. The
`levels of most observed ions were also similar between wild-type and
`R132H mutant IDH1 expressing cells (Fig. 1b), with no significant
`changes found in canonical tricarboxylic acid (TCA) cycle species
`(P . 0.05 for citrate, isocitrate, a-ketoglutarate, succinate, fumarate
`and malate). However, three species were significantly more abundant
`in R132H mutant IDH1 expressing cells (P , 0.001 for each). The
`mass of one of these ions matched precisely to 2HG (expected m/z
`147.0299; measured 147.0299). The other ions co-eluted with 2HG,
`and had masses consistent with the sodium adduct and a dehydrated
`form of 2HG. Subsequent injection of 2HG standard confirmed a
`retention time match to the biological peak, and that it forms all
`three of the observed ions during LC–MS ionization (data not shown).
`The cellular accumulation of 2HG was quantified by targeted
`
`1Agios Pharmaceuticals, Cambridge, Massachusetts 02139, USA. 2Department of Chemistry and Integrative Genomics, Princeton University, Princeton, New Jersey 08544, USA.
`3Department of Neurosurgery, UCLA Medical School, Los Angeles, California 90095, USA. 4Abramson Cancer Center, University of Pennsylvania, Philadelphia, Pennsylvania 19104,
`USA. 5Beth Israel Deaconess Medical Center and Harvard Medical School, Boston, Massachusetts 02215, USA. {Present address: Koch Institute for Integrative Cancer Research,
`Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA.
`
`
`
` ©2009
`
`Macmillan Publishers Limited. All rights reserved
`
`739
`
`Rigel Exhibit 1024
`Page 1 of 8
`
`

`

`ARTICLES
`
`NATURE | Vol 462 | 10 December 2009
`
`B
`
`A
`
`C
`
`107
`
`106
`
`105
`
`R132H signal intensity
`
`b
`
`a
`
`Blot:
`anti-IDH1
`Blot:
`anti-Myc
`
`ID H1(R132H)– M ycID H1– M ycParental
`
`
`
`
`
`IDH1–Myc
`
`IDH1
`endogenous
`
`1.0 × 106
`7.5 × 105
`5.0 × 105
`2.5 × 105
`0
`
`(a.u.)
`
`Peak intensity
`
`A
`
`149.50
`
`m/z
`
`146.50
`
`11.5
`
`OH
`
`HO
`
`Time (min)
`
`14.0
`
`2HG
`
`OH
`
`O
`O
`m/z 147.0299 (expected 147.0299, C5H7O5)
`
`104
`
`104
`
`107
`106
`105
`Wild-type signal intensity
`
`Parental
`
`IDH1–Myc
`
`IDH1(R132H)–Myc
`
`4,000
`
`3,200
`
`2,400
`
`1,600
`
`800
`
`d
`
`2HG (ng ml–1)
`
`120
`100
`80
`60
`40
`20
`0
`
`2HG (fold change)
`
`c
`
`ID H1– M ycParental
`
`ID H1(R132H)– M yc
`
`0
`
`0.5 h
`Figure 1 | Cells expressing human R132H IDH1 contain markedly elevated
`levels of 2HG. a, Western blots for Myc-tagged human isocitrate
`dehydrogenase 1 (IDH1–Myc) or R132H mutant (IDH1(R132H)–Myc) in
`stably transfected U87MG human glioblastoma cells. b, Metabolite profiles
`from cells expressing R132H IDH1 or wild-type IDH1 detected by LC–MS
`scanning for species between 110–1,000 m/z (M-H1). Red spots labelled A, B
`and C represent species assigned to 2HG, dehydro-2HG and 2HG-sodium
`adduct, respectively. Spectrometric details supporting the identification of
`
`triple-quadrupole LC–MS–MS analysis of cell extracts (Fig. 1c). The
`structure of 2HG is close to a-ketoglutarate, the product of the IDH1
`enzyme. Thus, the sole metabolite identified by untargeted metabolite
`profiling to be markedly altered by R132H mutant IDH1 expression
`was also implicated by its structure to be IDH1-related.
`The accumulation of 2HG was not restricted to cell extracts, as
`2HG was found to accumulate rapidly in the medium of cells expres-
`sing R132H mutant IDH1 (Fig. 1d). No appreciable 2HG could be
`found in the medium of wild-type cells or cells transfected with wild-
`type IDH1 (Fig. 1d). Isotope-labelling experiments on whole cells
`using uniformly labelled 13C-glutamine as a culture media nutrient
`demonstrated that the carbons in 2HG are derived from glutamine,
`with reasonably high overall pathway flux from glutamine through
`glutamate and a-ketoglutarate to 2HG. Moreover, labelling experi-
`ments did not demonstrate any other major alterations in central
`carbon metabolic flux in cells expressing R132H mutant 2HG (data
`not shown). The presence of a Myc epitope tag did not alter activity of
`R132H mutant IDH1. Despite being expressed at lower levels than
`the Myc-tagged R132H IDH1, cells transfected with untagged R132H
`IDH1 demonstrated a comparable increase in 2HG production
`(Supplementary Fig. 1). To determine whether 2HG production in
`cells expressing R132H mutant IDH1 is unique to U87MG cells, we
`stably expressed wild-type and R132H mutant IDH1 in wild-type
`IDH1-expressing LN-18 glioblastoma cells (Supplementary Fig.
`2a). Similar to results obtained with U87MG cells, the major differ-
`ence in metabolite levels observed in LN-18 cells expressing R132H
`mutant IDH1 was an increased level of 2HG (Supplementary Fig. 2b).
`
`Mutant IDH1 directly converts a-ketoglutarate to 2HG
`The R132H mutation has been reported to result in loss of function
`for enzyme activity3,7. However, in these studies only the NADP1-
`dependent oxidative decarboxylation of isocitrate to a-ketoglutarate
`was assessed. To understand how IDH1 activity is altered in cells by
`the presence of R132H mutant IDH1, we expressed increasing
`
`740
`
`4 h
`
`24 h
`
`species ‘A’ as 2HG are shown in the right panel. c, Cells expressing R132H
`IDH1 contain elevated levels of 2HG. Data were normalized by cell number
`and expressed as fold difference relative to parental values. Error bars depict
`one standard deviation (s.d.) from the mean of three independent
`experiments. d, Cells expressing R132H IDH1 display time-dependent
`accumulation of 2HG in cell culture media, normalization was as described
`in c. Errors bars depict one s.d. from the mean of four independent
`experiments.
`
`amounts of wild-type and R132H mutant IDH1 separately or in
`combination and assessed isocitrate-dependent NADPH production
`and a-ketoglutarate-dependent NADPH consumption in cell lysates.
`Consistent with published results, expression of R132H mutant
`IDH1 resulted in no measurable production of NADPH from isoci-
`trate, and isocitrate-dependent NADPH production increased with
`increasing amounts of wild-type enzyme (Supplementary Fig. 3a, b).
`The ability of the wild-type enzyme to generate NADPH was
`decreased slightly by co-expression of the R132H mutant IDH1.
`Opposite results were obtained, however, when NADPH consump-
`tion was measured in the presence of a-ketoglutarate. NADPH con-
`sumption by wild-type enzyme was not observed, whereas R132H
`mutant IDH1 expression resulted in a-ketoglutarate-dependent
`NADPH consumption (Supplementary Fig. 3c). Although the overall
`consumption of NADPH was slow, if anything co-expression of wild-
`type IDH1 with R132H mutant IDH1 facilitated the a-ketoglutarate-
`dependent consumption of NADPH. These findings demonstrate
`that in contrast to wild-type IDH1, R132H mutant IDH1 promotes
`an NADPH-dependent reduction of a-ketoglutarate. Furthermore,
`as this reduction was not inhibited by co-expression of wild-type
`IDH1, these data indicate that the novel activity of mutant IDH1
`can persist even in the presence of a wild-type IDH1 allele. In fact,
`it is possible that in the case of a heterodimer of wild-type and mutant
`IDH1, the a-ketoglutarate and NADPH produced locally by the wild-
`type subunit could be used as substrates for the mutant subunit,
`explaining the decrease in NADPH production observed in lysates
`when wild-type and mutant IDH1 are co-expressed.
`To understand how R132 mutations alter the enzymatic properties
`of IDH1, wild-type and R132H mutant IDH1 proteins were pro-
`duced and purified from Escherichia coli. When NADP1-dependent
`oxidative decarboxylation of isocitrate was measured using purified
`wild-type or R132H mutant IDH1 protein, it was confirmed that
`R132H mutation impairs the ability of IDH1 to catalyse this reac-
`tion3,7, as evident by the loss in binding affinity for both isocitrate and
`
`
`
` ©2009
`
`Macmillan Publishers Limited. All rights reserved
`
`Rigel Exhibit 1024
`Page 2 of 8
`
`

`

`NATURE | Vol 462 | 10 December 2009
`
`ARTICLES
`
`Table 1 | R132H mutation alters the enzymatic properties of IDH1
`Wild-type IDH1
`R132H IDH1
`Kinetic parameter of reaction
`
`O
`
`O
`
`HO
`
`OH
`
`OH
`
`2HG control
`
`Reductive rxn
`product
`
`0
`
`5
`HPLC elution time (min)
`
`10
`
`S(+)-2HG
`
`rxn + racemate
`
`2HG racemate
`
`R(–)-2HG
`
`O
`
`O
`
`HO
`
`OH
`
`OH
`
`rxn + R(–)-2HG
`R(–)-2HG std
`
`Reductive rxn
`
`100
`
`80
`
`60
`
`40
`
`20
`
`0
`
`150
`
`100
`
`50
`
`0
`
`a
`
`LC–MS–MS signal (normalized)
`
`b
`
`LC–MS–MS signal (normalized)
`
`10
`0
`HPLC elution time (min)
`
`20
`
`Figure 2 | R132H mutation in IDH1 results in production of R(2)-2HG.
`a, 2HG was identified as the reductive reaction product of recombinant
`human R132H mutant IDH1 using LC–MS as shown. b, The chirality of
`2HG produced by R132H mutant IDH1 was assessed as by diacetyl-L-tartaric
`anhydride derivatization and LC–MS analysis. Normalized LC–MS signal
`for the reductive reaction (rxn) product alone, an R(2)-2HG standard
`alone, and the two together (rxn 1 R(2)-2HG) are shown, as is the signal for
`a racemic mixture of R(2) and S(1) forms (2HG racemate) alone or with the
`reaction products (rxn 1 racemate).
`
`Two important features were noted by the change of R132 to
`histidine: the effect on conformation equilibrium and the reorganization
`of the active site. Located atop a b-sheet in the relatively rigid small
`domain, R132 acts as a gate-keeper residue and seems to orchestrate
`the hinge movement between the open and closed conformations. The
`guanidinium moiety of R132 swings from the open to the closed con-
`formation with a distance of nearly 8 A˚ . Substitution of histidine for
`arginine is likely to change the equilibrium in favour of the closed
`conformation that forms the catalytic cleft for cofactor and substrate
`to bind efficiently, which partly explains the high affinity for NADPH
`shown by the R132H mutant enzyme. This feature may be advantageous
`for the NADPH-dependent reduction of a-ketoglutarate to R(2)-2HG
`in an environment where NADPH concentrations are low. Second,
`closer examination of the catalytic pocket of the mutant IDH1 structure
`in comparison to the wild-type enzyme showed not only the expected
`loss of key salt-bridge interactions between the guanidinium of R132 and
`the a/b carboxylates of isocitrate, as well as changes in the network that
`coordinates the metal ion, but also an unexpected reorganization of the
`active site. Mutation to histidine resulted in a significant shift in position
`of the highly conserved residues Y139 from the A subunit and K2129
`from the B subunit (Fig. 3b, c), both of which are thought to be critical
`for catalysis by this enzyme family11. In particular, the hydroxyl moiety of
`Y139 now occupies the space of the b-carboxylate of isocitrate.
`The electron density in the active site was not sufficient to assign
`a-ketoglutarate and its orientation unambiguously. We have mod-
`elled the substrate based on the available electron density, taking into
`consideration the coordination between the carbonyl oxygen of
`a-ketoglutarate and the calcium ion as well as an orientation of
`a-ketoglutarate that would produce R(2)-2HG, the experimental
`product. The model required a significant repositioning of a-keto-
`glutarate compared to isocitrate, such that the distal carboxylate of
`
`741
`
`Oxidative (RNADPH)
`1 (mM)
`Km,NADP
`Km,isocitrate (mM)
`Km,MgCl2 (mM)
`Ki,a-ketoglutarate (mM)
`kcat (s21)
`Reductive (RNADP1)
`Km,NADPH (mM)
`Km,a-ketoglutarate (mM)
`kcat (s21)
`Kinetic parameters of oxidative and reductive reactions as measured for wild-type and R132H
`IDH1 enzymes are shown. Km (Michaelis constant) and kcat values for the reductive activity of
`the wild-type enzyme were unable to be determined as no measurable enzyme activity was
`detectable at any substrate concentration. Ki, inhibition constant.
`* No measurable enzymatic activity.
`
`49
`65
`29
`1.9 3 103
`4.4 3 104
`
`n/a*
`n/a
`n/a
`
`84
`370
`1.0 3 104
`24
`37.5
`
`0.44
`965
`1.0 3 103
`
`MgCl2 along with a 1,000-fold decrease in catalytic turnover (Table 1
`and Supplementary Fig. 4a). In contrast, when NADPH-dependent
`reduction of a-ketoglutarate was assessed using either wild-type or
`R132H mutant IDH1 protein, only R132H mutant could catalyse this
`reaction (Table 1 and Supplementary Fig. 4b). Part of this increased
`rate of a-ketoglutarate reduction results from an apparent increase in
`affinity for both the cofactor NADPH and substrate a-ketoglutarate
`in the R132H mutant IDH1 (Table 1). Taken together, these data
`demonstrate that whereas the R132H mutation leads to a loss of
`enzymatic function for oxidative decarboxylation of isocitrate, this
`mutation also results in a gain of enzyme function for the NADPH-
`dependent reduction of a-ketoglutarate.
`Reduction of the a-ketone in a-ketoglutarate can result in 2HG.
`To determine whether R132H mutant protein directly produced
`2HG from a-ketoglutarate we examined the product of the mutant
`IDH1 reaction using negative ion mode triple-quadrupole electro-
`spray LC–MS. These experiments confirmed that 2HG was the direct
`product of NADPH-dependent a-ketoglutarate reduction by the
`purified R132H mutant protein through comparison with known
`metabolite standards (Fig. 2a). Conversion of a-ketoglutarate to iso-
`citrate was not observed. To determine the chirality of the 2HG
`produced, we derivatized the products of the R132H reaction with
`diacetyl-L-tartaric anhydride, which allowed us to separate the (S)
`and (R) enantiomers of 2HG by simple reverse-phase LC and detect
`the products by tandem mass spectrometry9 (Fig. 2b). The peaks
`corresponding to the (S) and (R) isomers of 2HG were confirmed
`using racemic and R(2)-2HG standards. The reaction product from
`R132H co-eluted with the R(2)-2HG peak, demonstrating that the
`R(2) stereoisomer is the product produced from a-ketoglutarate by
`R132H mutant IDH1.
`To determine whether the altered enzyme properties resulting
`from the R132H mutation were shared by other R132 mutations
`found in human gliomas, recombinant R132C, R132L and R132S
`mutant IDH1 proteins were generated and the enzymatic properties
`assessed. Similar to R132H mutant protein, R132C, R132L and
`R132S mutations all result in a gain-of-function for NADPH-
`dependent reduction of a-ketoglutarate (Supplementary Fig. 4).
`Thus, in addition to impaired oxidative decarboxylation of isocitrate,
`one common feature shared among the IDH1 mutations found in
`human gliomas is the ability to catalyse direct NADPH-dependent
`reduction of a-ketoglutarate.
`
`X-ray structure reveals a distinct active site of mutant IDH1
`To define how R132 mutations alter the enzymatic properties of
`IDH1, the crystal structure of R132H mutant IDH1 bound to a-
`ketoglutarate, NADPH and Ca21 was solved at 2.1 A˚ resolution (see
`Supplementary Table 1 for crystallographic data and refinement sta-
`tistics). The overall quaternary structure of the homodimeric R132H
`mutant enzyme adopts the same closed catalytically competent
`conformation (shown as a monomer in Fig. 3a) that has been prev-
`iously described for the wild-type enzyme10.
`
`
`
` ©2009
`
`Macmillan Publishers Limited. All rights reserved
`
`Rigel Exhibit 1024
`Page 3 of 8
`
`

`

`ARTICLES
`
`NATURE | Vol 462 | 10 December 2009
`
`a
`
`b
`
`c
`
`Closed
`(R132H)
`
`vs
`
`Closed
`(WT)
`
`Closed
`(R132H)
`
`vs
`
`Open
`(WT)
`
`WT
`
`Arg 132
`
`AArg 100
`
`Tyr 139
`
`AAAAsn 96
`
`Ser 94
`
`
`
`Arg AA 109Arg 109
`
`Asp 275
`
`Ca2+
`
`Thr 77
`
`Thr 214′
`
`Lys 212′
`Asp 252′
`
`NADP+
`
`Ser 94
`
`Thr 77
`Thr 214′
`
`NADPH
`
`Tyr 139
`
`
`
`96AAAsnAAsn 9
`
`
`
`R132H
`
`His 132
`
`Arg rg 100
`
`
`
`Arg gArg 10
`
`9
`
`Asp 275
`
`Ca2+
`
`Lys 212′
`Asp 252′
`
`′
`
`′
`
`Figure 3 | Structural analysis of R132H mutant IDH1. a, On the left is shown
`an overlay of R132H mutant IDH1 (green) and wild-type IDH1 (grey)
`structures in the ‘closed’ conformation. On the right is shown an overlay of
`wild-type IDH1 (blue) structure in the ‘open’ conformation with mutant
`IDH1 (green) for comparison. b, Close-up comparison of the R132H IDH1
`active site (left) with a-ketoglutarate (yellow) and NADPH (grey) and the
`wild-type IDH1 active site (right) with isocitrate (yellow) and NADP (grey).
`
`Residues coming from the other monomer are denoted with a prime (9)
`symbol. In addition to the mutation at residue 132, the major changes are the
`positions of the catalytic residues Tyr 139 and Lys 2129. c, Wall-eyed stereo
`image showing the composite omit map for a-ketoglutarate, NADPH,
`calcium ion, His 132 and other key catalytic residues in the R132H mutant
`active site contoured at the 1s level.
`
`a-ketoglutarate now points upward to make new contacts with N96
`and S94. Overall, this single R132H mutation results in formation of
`a distinct active site compared to wild-type IDH1.
`
`2HG levels are elevated in human glioma samples
`Our data demonstrate that mutation of R132 can result in the ability
`of IDH1 to generate R(2)-2HG from a-ketoglutarate. To determine
`if 2HG production is characteristic of tumours harbouring muta-
`tions in IDH1, metabolites were extracted from human malignant
`gliomas that were either wild-type or mutant
`for IDH1 (see
`Supplementary Table 2 for summary of tumour characteristics). It
`has been suggested that a-ketoglutarate levels are decreased in cells
`transfected with mutant IDH17. We observe that the average a-keto-
`glutarate level from 12 tumour samples harbouring various R132
`mutations was slightly less than the average a-ketoglutarate level
`observed in 10 tumours which are wild type for IDH1. This difference
`
`in a-ketoglutarate was not statistically significant, and a range of
`a-ketoglutarate levels was observed in both wild-type and mutant
`tumours (Fig. 4). Similarly, a range of levels was observed for other
`proximal TCA cycle metabolites with no significant differences
`observed between wild-type IDH1 tumours and tumours with
`R132 IDH1 mutations. In contrast, increased 2HG levels were found
`in all tumours that contained an R132 IDH1 mutation (Fig. 4 and
`Supplementary Table 2). All R132 mutant IDH1 tumours examined
`had between 5 and 35 mmol of 2HG per gram of tumour, whereas
`tumours with wild-type IDH1 had over 100-fold less 2HG. This
`increase in 2HG in R132 mutant tumours was statistically significant
`(P , 0.0001). We confirmed that (R)-2HG was the isomer present in
`tumour samples (data not shown). Together these data establish that
`the novel enzymatic activity associated with R132 mutations in IDH1
`results in the production of 2HG in human brain tumours that har-
`bour these mutations.
`
`742
`
`
`
` ©2009
`
`Macmillan Publishers Limited. All rights reserved
`
`Rigel Exhibit 1024
`Page 4 of 8
`
`

`

`NATURE | Vol 462 | 10 December 2009
`
`ARTICLES
`
`catalytic activity, are not found. Finally, these findings have clinical
`implications in that they suggest that 2HG production will identify
`patients with IDH1 mutant brain tumours. This will be important for
`prognosis as patients with IDH1 mutations live longer than patients
`with gliomas characterized by other mutations5. In addition, patients
`with lower-grade gliomas may benefit by the therapeutic inhibition
`of 2HG production. Inhibition of 2HG production by mutant IDH1
`might slow or halt conversion of lower-grade glioma into lethal sec-
`ondary glioblastoma, changing the course of the disease.
`
`METHODS SUMMARY
`R132H, R132C, R132L and R132S mutations were introduced into human IDH1
`by standard molecular biology techniques. 293T and human glioma U87MG and
`LN-18 cell lines were transfected using standard techniques. Protein expression
`levels were determined by western blot. Metabolites were extracted from cul-
`tured cells and from tissue samples using 80% aqueous methanol (280 uC) as
`previously reported8. Metabolite levels in samples were determined by negative
`mode electrospray LC–MS. For untargeted profiling, extract components were
`resolved using reverse-phase high-performance liquid chromatography (HPLC)
`and metabolites were detected in ultra-high resolution mode (resolution
`,100,000) by stand alone orbitrap MS, collecting at one scan per second over
`an m/z range of 110–1,100. For targeted evaluation of 2HG, a-ketoglutarate and
`other TCA intermediates, extracts were resolved by reverse-phase HPLC system
`and metabolites detected by triple-quadrupole mass spectrometry, using
`multiple-reaction monitoring. Enzymatic activity in cell lysates was assessed
`by following a change in NADPH fluorescence over time in the presence of
`isocitrate and NADP, or a-ketoglutarate and NADPH. For enzyme assays using
`recombinant IDH1 enzyme, proteins were purified from E. coli using Ni affinity
`and size-exclusion chromatography. Enzymatic activity for recombinant IDH1
`protein was assessed by following a change in NADPH absorbance at 340 nm
`using a stop-flow spectrophotometer. Chirality of 2HG was determined as
`described previously9. For crystallography studies, purified R132H IDH1 was
`pre-incubated with NADPH, calcium chloride and a-ketoglutarate. Crystals
`were obtained at 20 uC by vapour diffusion equilibration using 3 ml drops mixed
`2:1 (protein:precipitant) against a well-solution of MES pH 6.5 and PEG 6000.
`Patient tumour samples were obtained after informed consent as part of a UCLA
`IRB-approved research protocol, collected by surgical resection, snap frozen in
`isopentane cooled by liquid nitrogen and stored at 280 uC. The IDH1 mutation
`status of each sample was determined as described previously3.
`
`Full Methods and any associated references are available in the online version of
`the paper at www.nature.com/nature.
`
`Received 15 July; accepted 29 October 2009.
`Published online 22 November 2009.
`
`α-
`Ketoglutarate Malate
`
`2HG
`
`Fumarate
`
`Succinate
`
`Isocitrate
`
`Mutant
`Wild type
`
`Mutant
`Wild type
`
`Mutant
`Wild type
`
`Mutant
`Wild type
`
`Mutant
`Wild type
`
`Mutant
`Wild type
`
`100.00
`
`10.00
`
`1.00
`
`0.10
`
`0.01
`
`Metabolite concentration (μmol g–1)
`
`Figure 4 | Human malignant gliomas containing R132 mutations in IDH1
`contain increased concentrations of 2HG. Human glioma samples obtained by
`surgical resection were snap frozen, genotyped to stratify as wild type (n 5 10)
`or carrying an R132 mutant allele (mutant) (n 5 12) and metabolites extracted
`for LC–MS analysis. Among the 12 mutant tumours, 10 carried a R132H
`mutation, one an R132S mutation, and one an R132G mutation. Each symbol
`represents the amount of the listed metabolite found in each tumour sample.
`Horizontal lines indicate the group sample means. The difference in 2HG
`observed between wild-type and R132 mutant IDH1 tumours was statistically
`significant by Student’s t-test (P , 0.0001). There were no statistically
`significant differences in a-ketoglutarate, malate, fumarate, succinate, or
`isocitrate levels between the wild-type and R132 mutant IDH1 tumours.
`
`Discussion
`2HG is known to accumulate in the inherited metabolic disorder
`2-hydroxyglutaric aciduria. This disease is caused by deficiency in
`the enzyme 2-hydroxyglutarate dehydrogenase, which converts 2HG
`to a-ketoglutarate12. Patients with 2-hydroxyglutarate dehydrogen-
`ase deficiencies accumulate 2HG in the brain as assessed by MRI and
`CSF analysis, develop leukoencephalopathy, and have an increased
`risk of developing brain tumours13–15. Furthermore, elevated brain
`levels of 2HG result in increased ROS levels16,17, potentially contrib-
`uting to an increased risk of cancer, and alterations in NADPH meta-
`bolism resulting from mutant IDH1 expression could further
`exacerbate this effect. The ability of 2HG to act as an NMDA (N-
`methyl-D-aspartate) receptor agonist may contribute to this effect16.
`2HG may also be toxic to cells by competitively inhibiting glutamate
`and/or a-ketoglutarate using enzymes. These include transaminases,
`which allow utilization of glutamate nitrogen for amino and nucleic
`acid biosynthesis, and a-ketoglutarate-dependent prolyl hydroxy-
`lases, such as those that regulate Hif1a levels. Alterations in Hif1a
`have been reported to result from mutant IDH1 protein expression7.
`Regardless of the mechanism, it seems likely that the gain-of-function
`ability of cells to produce 2HG as a result of R132 mutations in IDH1
`contributes to tumorigenesis. Patients with 2-hydroxyglutarate
`dehydrogenase deficiency have a high risk of central nervous system
`(CNS) malignancy15. The ability of mutant IDH1 to act directly on
`a-ketoglutarate may explain the prevalence of IDH1 mutations in
`tumours from CNS tissue, which are unique in their high level of
`glutamate uptake and its ready conversion to a-ketoglutarate in the
`cytosol18, thereby providing high levels of substrate for 2HG produc-
`tion. Myeloid cells also display a high ability to metabolize glutamine
`and recently R132 IDH1 mutations have also been described in a
`subset of acute myelogenous leukaemia (AML)19. The apparent co-
`dominance of the activity of mutant IDH1 with that of the wild-type
`enzyme is consistent with the genetics of the disease, in which only a
`single copy of the gene is mutated. As discussed above, the wild-type
`IDH1 could directly provide NADPH and a-ketoglutarate to the
`mutant enzyme. These data also demonstrate that mutation of
`R132 to histidine, serine, cysteine, glycine, or leucine shares a com-
`mon ability to catalyse the NADPH-dependent conversion of a-keto-
`glutarate to 2HG. These findings help clarify why mutations at other
`amino acid residues of IDH1, including other residues essential for
`
`1.
`
`5.
`
`6.
`
`Balss, J. et al. Analysis of the IDH1 codon 132 mutation in brain tumors. Acta
`Neuropathol. 116, 597–602 (2008).
`2. Watanabe, T., Nobusawa, S., Kleihues, P. & Ohgaki, H. IDH1 mutations are early
`events in the development of astrocytomas and oligodendrogliomas. Am. J.
`Pathol. 174, 1149–1153 (2009).
`3. Yan, H. et al. IDH1 and IDH2 mutations in gliomas. N. Engl. J. Med. 360, 765–773
`(2009).
`4. Hartmann, C. et al. Type and frequency of IDH1 and IDH2 mutations are related to
`astrocytic and oligodendroglial differentiation and age: a study of 1,010 diffuse
`gliomas. Acta Neuropathol. 118, 469–474 (2009).
`Parsons, D. W. et al. An integrated genomic analysis of human glioblastoma
`multiforme. Science 321, 1807–1812 (2008).
`Bleeker, F. E. et al. IDH1 mutations at residue p.R132 (IDH1(R132)) occur frequently
`in high-grade gliomas but not in other solid tumors. Hum. Mutat. 30, 7–11 (2009).
`7. Zhao, S. et al. Glioma-derived mutations in IDH1 dominantly inhibit IDH1 catalytic
`activity and induce HIF-1a. Science 324, 261–265 (2009).
`Lu, W., Kimball, E. & Rabinowitz, J. D. A high-performance liquid chromatography-
`tandem mass spectrometry method for quantitation of nitrogen-containing
`intracellular metabolites. J. Am. Soc. Mass Spectrom. 17, 37–50 (2006).
`Struys, E. A., Jansen, E. E., Verhoeven, N. M. & Jakobs, C. Measurement of urinary
`D- and L-2-hydroxyglutarate enantiomers by stable-isotope-dilution liquid
`chromatography-tandem mass spectrometry after derivatization with diacetyl-L-
`tartaric anhydride. Clin. Chem. 50, 1391–1395 (2004).
`10. Xu, X. et al. Structures of human cytosolic NADP-dependent isocitrate
`dehydrogenase reveal a novel self-regulatory mechanism of activity. J. Biol. Chem.
`279, 33946–33957 (2004).
`11. Aktas, D. F. & Cook, P. F. A lysine-tyrosine pair carries out acid-base chemistry in
`the metal ion-dependent pyridine dinucleotide-linked b-hydroxyacid oxidative
`decarboxylases. Biochemistry 48, 3565–3577 (2009).
`12. Struys, E. A. et al. Mutations in the D-2-hydroxyglutarate dehydrogenase gene
`cause D-2-hydroxyglutaric aciduria. Am. J. Hum. Genet. 76, 358–360 (2005).
`
`8.
`
`9.
`
`
`
` ©2009
`
`Macmillan Publishers Limited. All rights reserved
`
`743
`
`Rigel Exhibit 1024
`Page 5 of 8
`
`

`

`ARTICLES
`
`NATURE | Vol 462 | 10 December 2009
`
`13. Ko¨lker, S., Mayatepek, E. & Hoffmann, G. F. White matter disease in cerebral
`organic acid disorders: clinical implications and suggested pathomechanisms.
`Neuropediatrics 33, 225–231 (2002).
`14. Wajner, M., Latini, A., Wyse, A. T. & Dutra-Filho, C. S. The role of oxidative
`damage in the neuropathology of organic acidurias: insights from animal studies.
`J. Inherit. Metab. Dis. 27, 427–448 (2004).
`15. Aghili, M., Zahedi, F. & Rafiee, E. Hydroxyglutaric aciduria and malignant brain
`tumor: a case report and literature review. J. Neurooncol. 91, 233–236 (2009).
`16. Kolker, S. et al. NMDA receptor activation and respiratory chain complex V
`inhibition contribute to neurodegeneration in D-2-hydroxyglutaric aciduria. Eur. J.
`Neurosci. 16, 21–28 (2002).
`17. Latini, A. et al. D-2-hydroxyglutaric acid induces oxidati

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket