throbber
ARTICLE
`
`cros
`
`Molecular mechanisms of isocitrate dehydrogenase
`1 (IDH1) mutations identified in tumors: The role of size
`and hydrophobicity at residue 132 on catalytic efficiency
`
`Received for publication, January 10, 2017, and in revised form, March 16, 2017 Published, Papers in Press, March 22, 2017, DOI 10.1074/jbc.M117.776179
`Diego Avellaneda Matteo‡1, Adam J. Grunseth‡1, Eric R. Gonzalez‡, Stacy L. Anselmo‡, Madison A. Kennedy‡,
`Precious Moman‡, David A. Scott§, An Hoang‡, and Christal D. Sohl‡2
`From the ‡Department of Chemistry and Biochemistry, San Diego State University, San Diego, California 92182 and the §Sanford
`Burnham Prebys Medical Discovery Institute, La Jolla, California 92037
`
`Edited by John M. Denu
`
`Isocitrate dehydrogenase 1 (IDH1) catalyzes the reversible
`NADPⴙ-dependent conversion of isocitrate (ICT) to ␣-ketogl-
`utarate (␣KG) in the cytosol and peroxisomes. Mutations in
`IDH1 have been implicated in >80% of lower grade gliomas and
`secondary glioblastomas and primarily affect residue 132, which
`helps coordinate substrate binding. However, other mutations
`found in the active site have also been identified in tumors.
`IDH1 mutations typically result in a loss of catalytic activity, but
`many also can catalyze a new reaction, the NADPH-dependent
`reduction of ␣KG to D-2-hydroxyglutarate (D2HG). D2HG is
`a proposed oncometabolite that can competitively inhibit
`␣KG-dependent enzymes. Some kinetic parameters have been
`reported for several IDH1 mutations, and there is evidence that
`mutant IDH1 enzymes vary widely in their ability to produce
`D2HG. We report that most IDH1 mutations identified in
`tumors are severely deficient in catalyzing the normal oxidation
`reaction, but that D2HG production efficiency varies among
`mutant enzymes up to ⬃640-fold. Common IDH1 mutations
`have moderate catalytic efficiencies for D2HG production,
`whereas rarer mutations exhibit either very low or very high
`efficiencies. We then designed a series of experimental IDH1
`mutants to understand the features that support D2HG produc-
`tion. We show that this new catalytic activity observed in tumors
`is supported by mutations at residue 132 that have a smaller van
`der Waals volume and are more hydrophobic. We report that
`one mutation can support both the normal and neomorphic
`reactions. These studies illuminate catalytic features of muta-
`tions found in the majority of patients with lower grade gliomas.
`
`Metabolic changes in tumors have been described for nearly
`a century (1–3), but only relatively recently have enzymes
`
`This work was supported by National Institute of Health Grants K99 CA187594
`(to C. D. S.), R00 CA187594 (to C. D. S.), 5T34 GM008303 (to E. R. G. and
`M. A. K.), and P30 CA030199 (to D. A. S.), a Summer Undergraduate
`Research Program Grant from San Diego State University (to M. A. K.), and
`San Diego State University startup funds (to C. D. S.). The authors declare
`that they have no conflicts of interest with the contents of this article. The
`content is solely the responsibility of the authors and does not necessarily
`represent the official views of the National Institutes of Health.
`This article contains supplemental Figs. S1–S6.
`1 Both authors contributed equally to this work.
`2 To whom correspondence should be addressed: CSL 328, MC1030, 5500
`Campanile Dr., San Diego, CA 92182. Tel.: 619-594-2053; Fax: 619-594-
`4634; E-mail: csohl@mail.sdsu.edu.
`
`involved in metabolic processes been established as tumor sup-
`pressors or oncoproteins. One of the more striking examples of
`metabolic enzymes playing a role in tumorigenesis includes
`isocitrate dehydrogenase 1 and 2 (IDH1 and IDH2).3 These
`homodimeric enzymes are responsible for the reversible
`NADP⫹- and Mg2⫹-dependent conversion of ICT to ␣KG
`(Fig. 1A) in the cytosol and peroxisomes (IDH1), or mito-
`chondria (IDH2). IDH3 is responsible for the same reaction
`within the context of the TCA cycle, although the oxidative
`decarboxylation catalyzed by this enzyme is non-reversible
`and NAD⫹-dependent.
`Mutations in IDH1 and IDH2 were identified in glioblastoma
`multiforme in a large sequencing effort (4), and soon ⬎80% of
`adult grade II/III gliomas and secondary glioblastomas were
`found to have IDH1 mutations, commonly R132H or R132C
`IDH1 (5, 6) (reviewed in Refs. 7–9). Subsequently ⬃10–20% of
`acute myeloid leukemias were shown to have primarily IDH2
`mutations, typically R140Q or R172K IDH2 (10). Early mecha-
`nisms of tumorigenesis focused on deficient conversion of ICT
`to ␣KG (11), suggesting that IDH serves as a tumor suppressor,
`in part through altering levels of hypoxia-inducible transcrip-
`tion factor-1␣ (12). However, IDH1 and IDH2 mutations
`appeared heterozygously in tumors, an unusual feature of a
`tumor suppressor. In landmark studies (13–15), the most com-
`mon IDH1 and IDH2 mutations were shown to catalyze a neo-
`morphic reaction: the Mg2⫹- and NADPH-dependent reduc-
`tion of ␣KG to D2HG (Fig. 1B). This suggested IDH1 and IDH2
`likely encode for oncoproteins. D2HG is proposed to be an
`oncometabolite; it competitively inhibits ␣KG-dependent en-
`zymes including the TET family of 5-methylcytosine hydroxy-
`lases and the JmjC family of histone lysine demethylases,
`resulting in cell de-differentiation (16, 17). Indeed, cancer
`patients with IDH mutations display hypermethylated pheno-
`types (18–20) resulting from D2HG-mediated inhibition of
`histone and DNA demethylation. The proposed oncometabo-
`lite D2HG alone can recapitulate tumorigenic phenotypes in
`cancer models (21, 22), but studies measuring global metabo-
`lomics changes between mutant IDH1 expression and D2HG
`treatment show some differences (23, 24), indicating loss of the
`
`3 The abbreviations used are: IDH, isocitrate dehydrogenase; ICT, isocitrate;
`␣KG, ␣-ketoglutarate; D2HG, D-2-hydroxyglutarate; AML, acute myeloid
`leukemia.
`
`J. Biol. Chem. (2017) 292(19) 7971–7983 7971
`
`© 2017 by The American Society for Biochemistry and Molecular Biology, Inc. Published in the U.S.A.
`
`Rigel Exhibit 1051
`Page 1 of 13
`
`

`

`Catalytic efficiency of IDH1 mutants
`
`Figure 1. WT and mutant IDH1 catalytic activities. Shown are the: A, nor-
`mal oxidative decarboxylation, and B, the neomorphic reduction.
`
`normal reaction and/or altered NADPH levels may also play an
`important role. Development of selected targeted therapy
`against oncogenic IDH1 and IDH2 mutations is underway and
`results are promising (8).
`A point mutation conferring a new catalytic activity suggests
`important mechanistic features, and many have used kinetics
`and structural methods to explore mutant IDH1 and IDH2
`activity (13, 25–30). There is evidence that IDH1 mutations
`produce varying concentrations of D2HG (30), with somewhat
`subtle but interesting alterations in conformational changes as
`shown in crystal structures of R132H IDH1 (26, 27, 29). In gen-
`eral, reported kinetic parameters of IDH1 mutants explored to
`date vary widely, making comparisons difficult. Interestingly,
`some mutations identified in tumors do not appear to generate
`D2HG (31, 32), suggesting that loss of the normal reaction itself
`has important consequences, or perhaps that they are simply
`passenger mutations.
`Here we report a thorough catalytic study of a wide spectrum
`of IDH1 mutations, including many identified in tumors and
`several mutants designed to clarify catalytic features. We show
`that IDH1 mutants vary widely in catalytic efficiency, with more
`polar and larger residues at position 132 supporting the normal
`reaction, and more hydrophobic and smaller residues driving
`the neomorphic reaction. These findings provide significant
`insight into the types of mutations that may be accommodated
`at residue 132 for efficient D2HG production. By determining
`the catalytic features of IDH1 mutations, we reveal features of
`driver mutations present in the majority of patients with lower
`grade gliomas and secondary glioblastomas.
`
`Results
`Structural modeling and thermal stability of IDH1 mutations
`For the mutations explored in this work, only structures of
`WT and R132H IDH1 in complex with both substrates (ICT
`and NADP⫹, or ␣KG and NADP⫹) have been reported (13, 27,
`29, 33), although a recent high resolution cryo-EM structure
`shows R132C IDH1 in complex with NADPH (34). To help
`inform the structural consequences of R132C, R132G, R100Q,
`A134D, and H133Q IDH1, these mutants were modeled in pre-
`viously solved structures of R132H IDH1 in complex with ␣KG,
`NADP⫹, and Ca2⫹ (Protein Data Bank (PDB) 4KZO (27)) and
`WT IDH1 in complex with ICT, NADP⫹, and Ca2⫹ (PDB 1T0L
`(33)) using the geometry minimization package in Phenix (35).
`These models were then aligned to the original structures using
`PyMOL (36) (Fig. 2). In both models, minimal global changes
`
`7972 J. Biol. Chem. (2017) 292(19) 7971–7983
`
`Figure 2. Structural modeling of IDH1 mutations identified in tumors. A,
`the structure of WT IDH1 complexed with ICT, NADP⫹, and Ca2⫹ (PDB 1T0L
`(33)) and B, R132H IDH1 complexed with ␣KG, NADP⫹, and Ca2⫹ (PDB 4KZO
`(27)) were used to model additional mutations. In both panels, WT IDH1 is
`shown in green, A134D in cyan, H133Q in black, R100Q in dark blue, R132H in
`orange, R132C in yellow, and R132G in gray. Substrates and residues that are
`mutated are highlighted in stick format, as well as catalytic residue Tyr-139.
`Ca2⫹ is shown as a sphere. Ligand restraint generation and optimization of
`provided cif files were generated using eLBOW in the Phenix software suite
`(35), and mutations were made using Coot (54). Geometry Minimization (Phe-
`nix software suite) (35) was used to regularize geometries of the models, with
`500 iterations and 5 macro cycles.
`
`were identified, consistent with previous structural work on
`R132H IDH1. The presence of the C3 carboxylate in ICT
`requires some adjustment of the catalytic residue Tyr-139,
`whereas less movement of this residue is seen in the models
`containing ␣KG.
`To explore the mechanistic features of these IDH1 mutations
`identified in tumors, cDNA constructs were generated for het-
`erologous expression and purification in Escherichia coli. WT,
`R132H, R132C, R132G, R100Q, A134D, and H133Q IDH1
`homodimers were expressed and purified to ⬎95% purity
`
`Rigel Exhibit 1051
`Page 2 of 13
`
`

`

`(supplemental Fig. S1A). Thermal stability was assessed for
`each enzyme using circular dichroism to perform thermal shift
`assays. The melting temperature (Tm) of each of the IDH1
`mutants varied little compared with WT IDH1 (supplemental
`Fig. S2A). R132C IDH1 had the lowest Tm (46.8 °C), but this
`represents only a 5% change from WT IDH1 (Tm ⫽ 49.1 °C).
`R100Q IDH1 had the highest Tm (51.9 °C), again signifying only
`a 5% change from WT IDH1.
`
`Efficiency of reactions catalyzed by IDH1 mutants found in
`tumors
`The most common IDH1 mutations found in gliomas are
`R132H followed by R132C (37). R132G, which has been identi-
`fied with higher frequency in chondrosarcomas (38, 39), is less
`frequently seen in gliomas and is a known D2HG producer (30).
`R100Q IDH1, a long predicted D2HG producer based on the
`R140Q IDH2 mutation affecting the identical residue, is rela-
`tively rare (40). A134D and H133Q IDH1 are rare mutations
`found in thyroid cancers and are predicted to only be deficient
`in the normal reaction (31, 40). Steady-state kinetic assays were
`used to determine the catalytic efficiency of the conversion of
`ICT to ␣KG (normal reaction) by monitoring the production of
`NADPH at A340 nm, or ␣KG to D2HG (neomorphic reaction) by
`monitoring the consumption of NADPH, at both 21 and 37 °C.
`All mutants were deficient in the normal reaction, ranging from
`a relatively minor 3.5-fold loss of catalytic efficiency (kcat/Km)
`for H133Q IDH1, whereas the other mutations exhibited more
`severe ⬃300- to 1,340-fold losses in efficiency (Fig. 3, supple-
`mental Fig. S3, Table 1). The observed changes in catalytic effi-
`ciency are driven both by decreases in kcat and increases in Km.
`Mutants varied widely in their relative catalytic efficiency of
`D2HG production (Fig. 4, supplemental Fig. S4, Table 2). Only
`rate-saturating concentrations of substrates generated rates
`of D2HG production above the signal-to-noise threshold for
`A134D IDH1 and H133Q IDH1. Thus only upper limits of kcat
`values are reported as kobs (Fig. 4, B and D) because Km values
`could not be obtained. R132G IDH1 is the most efficient pro-
`ducer of D2HG (⬃125-fold more efficient than WT IDH1),
`driven primarily by low Km values but also by a high kcat. R132C
`and R132H IDH1 are ranked next in catalytic efficiency, with a
`low Km value reported for R132C IDH1 (Table 2). This suggests
`that production of D2HG in tumors by R132G and R132C IDH1
`may be more significant than R132H IDH1 when cytosolic con-
`centration of ␣KG is considered. This trend is supported by
`D2HG measurements in glioma tissue (30). Due to its high Km,
`R100Q IDH1 was one of the least efficient producers of D2HG
`(Table 2).
`
`GC/MS analysis confirms D2HG production by IDH1 tumor
`mutants
`Although the normal reaction is reversible (Fig. 1A), a lower
`pH and source of CO2 (typically NaHCO3) are required to favor
`the reverse reaction in vitro, and work by Leonardi et al. (25)
`have shown that the reverse reaction is deficient in IDH1
`mutants. Regardless, we desired to confirm that the less well
`characterized IDH1 mutants, namely R100Q and R132G, favor
`D2HG production over ICT when incubated with ␣KG and
`NADPH. R132H and R132C IDH1 are well established to pro-
`
`Catalytic efficiency of IDH1 mutants
`
`duce D2HG (first reported in Ref. 13, but confirmed by many
`groups). Gas chromatography/mass spectrometry (GC/MS)
`was used to identify and quantify the amount of D2HG as well
`as ␣KG and ICT (not shown) produced in these incubations
`(Fig. 5) (20). R132G IDH1 gave robust production of D2HG
`consistent with kinetic data, whereas an incubation with R100Q
`IDH1 showed D2HG production levels near the lower limit of
`detection, again consistent with kinetic findings (Figs. 4 and 5).
`Levels of ICT for both mutants were ⬍0.1 nmol, based on limits
`of detection. This indicates that NADPH oxidation in the pres-
`ence of ␣KG is preferably coupled to D2HG production under
`these reaction conditions. This also supports previous findings
`that both mutants generate D2HG in in vitro assays (30). These
`experiments do not necessarily indicate that the reverse of the
`normal reaction is ablated, however, as pH ⬍ 7 and CO2 are
`required for this reaction in vitro (25).
`
`Generation of IDH1 mutants engineered to explore
`mechanistic features of D2HG production
`In addition to R132H/R132C/R132G IDH1, several rarer
`IDH1 mutations in gliomas have been identified, including
`R132S/R132L/R132V IDH1 (5, 6, 41, 42). In vitro kinetic assays
`have shown that R132S and R132L IDH1 catalyze production of
`D2HG at rates similar to the more common R132H and R132C
`IDH1 mutations (13, 25). Similarly, ectopic expression of R132S
`and R132L IDH1 in HEK293T cells indicate D2HG production
`levels are comparable with cell lines expressing R132C and
`R132H IDH1 (30). R132H/R132C/R132G/R132S/R132L/
`R132V IDH1 all vary in the degree of hydrophobicity at residue
`132, and all have a smaller van der Waals volume than the
`wild-type arginine. Although these clues illuminate interesting
`mechanistic characteristics of R132H IDH1, the features that
`allow IDH1 mutants to generate D2HG with varying catalytic
`efficiency are not fully clear.
`We designed several IDH1 mutations to serve as tools to
`probe the limits of hydrophobicity (43) and van der Waals vol-
`ume (44) at residue 132 that support D2HG production. R132A
`IDH1 is truly an engineered mutation, as to our knowledge it
`has not been identified in tumors. This residue serves as an
`example of a more hydrophobic and smaller residue at position
`132, similar to R132G IDH1. R132A IDH1 has been shown to be
`deficient in the normal reaction (29), but its ability to catalyze
`the neomorphic reaction has not yet been explored. R132N
`IDH1 also has not been identified in tumors to date. Asparagine
`has a much smaller van der Waals volume than arginine,
`although the ranked polarities of these two amino acids are
`similar. R132Q IDH1 plays an important role in driving chon-
`drosarcomas and a small number of gliomas, and mouse
`mR132Q IDH1 generates D2HG about 20-fold more efficiently
`than human R132H IDH1 in vitro (39, 45). This mutation has
`the most similar ranking in hydrophobicity as compared with
`WT, but a smaller van der Waals volume. R132K IDH1 is ho-
`mologous to R172K IDH2, one of the most common D2HG-
`producing mutations seen in acute myeloid leukemia (11, 41).
`However, R132K IDH1 has not been reported in tumors, and
`the activity of this enzyme has not been assessed. R132K IDH1
`is most comparable with WT IDH1 when considering both van
`der Waals volume and polarity ranking of residue 132. Finally,
`
`J. Biol. Chem. (2017) 292(19) 7971–7983 7973
`
`Rigel Exhibit 1051
`Page 3 of 13
`
`

`

`Catalytic efficiency of IDH1 mutants
`
`Figure 3. Concentration dependence of the ICT concentration on the observed rate of NADPH production in the normal reaction (37 °C). The deter-
`mined kobs values were obtained from two different enzyme preparations to ensure reproducibility. The kobs values resulting from each of the two enzyme
`preparations are distinguished by using either a circle or an ⫻ in the plots. The observed rate constants (kobs) were calculated from the linear range of the slopes
`of plots of concentration versus time using GraphPad Prism software (GraphPad, San Diego, CA). These kobs values were then fit to a hyperbolic equation to
`generate kcat and Km values, and the standard error listed in Table 1 results from the deviance from these hyperbolic fits is indicated. The determined kobs values
`were obtained from two different enzyme preparations to ensure reproducibility. Results from assays at 21 °C are shown in supplemental Fig. S3. A, WT IDH1.
`B, H133Q IDH1. C, A134D IDH1. D, R100Q IDH1. E, R132H IDH1. F, R132C IDH1. G, R132G IDH1.
`R132H IDH1 in complex with ␣KG, NADP⫹, and Ca2⫹ (PDB
`R132W IDH1 is another example of an engineered mutation in
`that it has not been identified in tumors. It was selected to
`code 4KZO (27)) using the geometry minimization package in
`represent the most extreme case of a large van der Waals vol-
`Phenix (35) followed by alignment in PyMOL (36) (Fig. 6). Few
`ume coupled with high hydrophobicity.
`changes are observed globally or within the active site. The size
`of the amino acid at position 132 does necessitate some local
`adjustments to avoid steric hindrance, but overall, changes in
`the models are minimal.
`All five IDH1 mutants were successfully heterologously
`expressed and purified to ⬎95% purity (supplemental Fig. S1B).
`
`Structural modeling and thermal stability of engineered IDH1
`mutations
`Because no crystal structures of these mutants are currently
`available, each was modeled in a previously solved structure of
`
`7974 J. Biol. Chem. (2017) 292(19) 7971–7983
`
`Rigel Exhibit 1051
`Page 4 of 13
`
`

`

`Catalytic efficiency of IDH1 mutants
`
`Thermal stability was assessed using circular dichroism in ther-
`mal shift assays. Again, minimal changes were observed in Tm
`(supplemental Fig. S2B). R132K IDH1 had the highest Tm
`(49.8 °C), which varied from WT IDH1 by only 2%.
`
`Kinetic analysis of engineered IDH1 mutants
`The catalytic efficiency of the normal reaction was measured
`for all mutants, and efficiencies were plotted against relative
`hydrophobicity according to Monera et al. (43) (Fig. 7A), and
`against van der Waals volume (44) (Fig. 7B). All mutants were
`significantly deficient in converting ICT to ␣KG, driven both by
`a decrease in kcat as well as an increase in Km (Table 3, supple-
`mental Fig. S5). Two IDH1 mutations maintained moderate
`oxidative decarboxylation activity; R132Q and R132K IDH1
`had 33- and 56-fold losses of ␣KG production efficiency relative
`to WT IDH1, respectively. All other mutations had ⱖ220-fold
`decreases in catalytic efficiency.
`IDH1 mutants were also incubated with ␣KG and NADPH to
`measure presumptive D2HG production efficiency (Fig. 7,
`Table 4, supplemental Fig. S6). R132Q IDH1 was the most effi-
`cient D2HG producer of the mutants explored in this work,
`with 4-fold higher efficiency than the next most efficient
`mutant, R132G IDH1. There was a notable decrease in effi-
`ciency in all other mutants, with R132A IDH1 having similar
`catalytic efficiencies as R132G/R132C/R132H IDH1. R132N/
`R132K/R132W and WT IDH1 were all very poor at producing
`D2HG. The severely deficient catalytic efficiency seen in R132N
`IDH1 was primarily driven by a very high Km value (Table 4).
`This suggests that like R100Q IDH1, D2HG production by
`R132N IDH1 may not be physiologically relevant when the
`cytosolic concentration of ␣KG is considered. Relative efficien-
`cies of the other mutants were driven both by changes in kcat
`and Km, with a low Km value driving R132A IDH1 production
`(Table 4).
`
`GC/MS analysis confirms D2HG production by engineered
`IDH1 mutants
`To confirm that an incubation of the engineered IDH1
`mutants with ␣KG and NADPH favors D2HG production
`rather than the reverse of the normal reaction (i.e. ICT produc-
`tion), GC/MS was used to quantify levels of D2HG, ICT, and
`␣KG of the engineered IDH1 mutants. Measured amounts of
`D2HG were as expected under the incubation lengths at exper-
`imentally measured kinetic efficiency. R132Q and R132G IDH1
`generated the highest levels of D2HG (Fig. 5), followed by
`R132A IDH1. Again, levels of ICT were difficult to measure due
`to their very low concentrations (⬍0.1 nmol, based on limits of
`detection). This suggests that NADPH oxidation is coupled
`primarily to D2HG production, rather than ICT, under these
`experimental conditions.
`
`Discussion
`Here we report the first in-depth, simultaneous catalytic
`characterization of 11 IDH1 mutations and WT IDH1, includ-
`ing mutations identified in tumors (R132H/R132C/R132G/
`R132Q, R100Q, A134D, and H133Q IDH1) and additional
`mutations (R132A/R132K/R132N/R132W IDH1) designed to
`measure the effects of hydrophobicity (43) and van der Waals
`
`J. Biol. Chem. (2017) 292(19) 7971–7983 7975
`
`0.28⫾0.05
`0.30⫾0.05
`0.020⫾0.003
`0.16⫾0.02
`0.074⫾0.003
`35⫾9
`7.3⫾1.3⫻102
`mM⫺1s⫺1
`
`0.14⫾0.02
`0.58⫾0.08
`1.0⫾0.9
`0.070⫾0.006
`0.7⫾0.1
`0.101⫾0.008
`0.03⫾0.01
`
`mM
`
`3.6⫾0.6
`5.3⫾0.8
`6⫾1
`9⫾1
`2.7⫾0.3
`0.28⫾0.07
`0.015⫾0.003
`
`1.0⫾0.06
`1.61⫾0.08
`0.120⫾0.006
`1.40⫾0.06
`0.200⫾0.008
`9.4⫾0.6
`11.0⫾0.4
`s⫺1
`
`1.3⫾0.2
`0.54⫾0.05
`0.57⫾0.08
`0.7⫾0.2
`0.29⫾0.08
`1.1⫾0.2⫻102
`3.9⫾0.4⫻102
`mM⫺1s⫺1
`
`0.067⫾0.007
`0.75⫾0.07
`1.6⫾0.5
`0.18⫾0.02
`1.2⫾0.3
`0.16⫾0.02
`0.08⫾0.03
`
`mM
`
`7⫾1
`8.2⫾0.8
`4.2⫾0.6
`8⫾2
`8⫾2
`0.40⫾0.08
`0.22⫾0.02
`
`9.3⫾0.6
`4.4⫾0.1
`2.4⫾0.1
`5.6⫾0.4
`2.3⫾0.2
`45⫾2
`85⫾4
`s⫺1
`
`48(Gly)
`86(Cys)
`118(His)
`148(Arg)
`148(Arg)
`148(Arg)
`148(Arg)
`
`0(Gly)
`49(Cys)
`8(His)
`⫺14(Arg)
`⫺14(Arg)
`⫺14(Arg)
`⫺14(R)
`
`R132G
`R132C
`R132H
`R100Q
`A134D
`H133Q
`WT
`
`bFromRef.44.
`aFromRef.43.
`
`(kcat/Km,ICT,21°C)
`Efficiency,mMⴚ1sⴚ1
`
`(21°C)
`Km,ICT
`
`(21°C)
`kcat
`
`(kcat/Km,ICT,37°C)
`
`Efficiency,
`
`(37°C)
`Km,NADPⴙ
`
`(37°C)
`Km,ICT
`
`(37°C)
`kcat
`
`vanderWaalsvolume
`
`residue132,Å3b
`ofsidechainat
`
`(21°C)
`Km,NADPⴙ
`Valuesresultfromfitsofkineticdatausingtwodifferentenzymepreparations.Thestandarderrorisdeterminedfromthedeviancefromthesehyperbolicfits(Fig.3,supplementalFig.S3).
`Kineticparametersforthenormalreaction,conversionofICTto␣KG,catalyzedbyIDH1
`Table1
`
`hydrophobicityof
`
`residue132a
`
`Relative
`
`IDH1
`
`Rigel Exhibit 1051
`Page 5 of 13
`
`

`

`Catalytic efficiency of IDH1 mutants
`
`Figure 4. Concentration dependence of ␣KG concentration on the observed rate of NADPH depletion in the neomorphic reaction (37 °C). The
`determined kobs values were obtained from two different enzyme preparations to ensure reproducibility. The kobs values resulting from each of the two enzyme
`preparations are distinguished by using either a circle or an ⫻ in the plots. The observed rate constants (kobs) were calculated from the linear range of the slopes
`of plots of concentration versus time using GraphPad Prism software (GraphPad). These kobs values were then fit to a hyperbolic equation to generate kcat and
`Km values, and the S.E. results from the deviance from these hyperbolic fits is indicated. Km values and efficiency are in terms of [␣KG]. Due to limits of detection,
`Km values could not be obtained for low efficiency IDH1 enzymes because only saturating kobs rates could be detected. In this case, kobs rates are reported,
`which approximate kcat rates. Results from assays at 21 °C are shown in supplemental Fig. S4. A, WT IDH1. B, H133Q IDH1. D, R100Q IDH1. E, R132H IDH1. F, R132C
`IDH1. G, R132G IDH1.
`volume (44) at residue 132 on catalytic efficiency. To date, a
`relatively wide range of catalytic rates have been reported for
`WT and mutant IDH1. For WT IDH1, kcat values for the normal
`reaction typically range from ⬃9 to 12 s⫺1 (26, 27, 29) at ambi-
`ent temperature. This is in good agreement with our findings
`(Table 1), although there are reports of much higher rates (13,
`15). Km values for both ICT and NADP⫹ typically range from 5
`to 65 ␮M at ambient temperature (13, 15, 25–27, 29), again
`
`consistent with our values (Table 1). Plasma levels of ␣KG have
`been reported to be ⬃23 ␮M (46), which is in line with measured
`Km values.
`H133Q IDH1 displays a relatively minor change in catalytic
`efficiency for ␣KG production compared with WT IDH1, and
`D2HG production was extremely slow, indicating that this may
`be a passenger mutation (Tables 1 and 2). R100Q IDH1 shows
`drastic increases in Km values both for ICT and ␣KG, resulting
`
`7976 J. Biol. Chem. (2017) 292(19) 7971–7983
`
`Rigel Exhibit 1051
`Page 6 of 13
`
`

`

`Catalytic efficiency of IDH1 mutants
`
`Figure 5. Absolute quantitation of D2HG present in an incubation of
`IDH1 mutants with ␣KG and NADPH. Measurements are reported as a cal-
`culated mean ⫾ S.E. Only trace amounts of ICT (⬍0.1 nmol, based on limits of
`detection) were generated under these experimental conditions, indicating
`that NADPH oxidation was coupled to D2HG production, rather than ICT
`production.
`
`Figure 6. Structural models of experimental IDH1 mutants. The structure
`of R132H IDH1 complexed with ␣KG, NADP⫹, and Ca2⫹ (PDB 4KZO (27)) was
`used to model mutations of the tool IDH1 mutations. R132H IDH1 is shown in
`orange, R132Q in magenta, R132N in cyan, R132A in dark blue, R132K in black,
`and R132W in purple. Substrates and residues that are mutated are high-
`lighted in stick format, as well as catalytic residue Tyr-139. Ca2⫹ is shown as a
`sphere. Ligand restraint generation and optimization of provided cif files were
`generated using eLBOW in the Phenix software suite (35), and mutations
`were made using Coot (54). Geometry Minimization (Phenix software suite)
`(35) was used to regularize geometries of the models, with 500 iterations and
`5 macro cycles.
`
`in significant decreases in catalytic efficiency for both reactions
`studied (Tables 1 and 2). This was surprising because the ho-
`mologous mutation in IDH2, R140Q, is the most common IDH
`mutation found in AML (47). However, kinetic characteriza-
`tions of IDH2 mutants are limited (28, 48). Kinetic and struc-
`
`J. Biol. Chem. (2017) 292(19) 7971–7983 7977
`
`mM⫺1s⫺1
`
`1.5⫾0.2
`2.3⫾0.4
`0.24⫾0.08
`0.013⫾0.001
`ND
`ND
`ND
`
`(kcat/Km,␣KG,21°C)
`
`Efficiency
`
`ⱕ0.025
`ⱕ0.025
`ⱕ0.005
`ⱕ0.0025
`ND
`ND
`ND
`
`(21°C)
`Km,NADPH
`
`mM
`
`0.30⫾0.05
`0.36⫾0.06
`1.8⫾0.6
`10⫾1
`ND
`ND
`NDc
`
`(21°C)
`Km,␣KG
`
`0.45⫾0.02
`0.84⫾0.03
`0.43⫾0.04
`0.128⫾0.006
`ⱕ0.020
`ⱕ0.034
`ⱕ0.017
`s⫺1
`
`sⴚ1(21°C)
`
`kcat,
`
`mM⫺1s⫺1
`
`5⫾1
`4.4⫾0.6
`3.8⫾0.9
`0.028⫾0.005
`ND
`ND
`0.04⫾0.02
`
`⬍0.025
`0.010⫾0.009
`ⱕ0.025
`0.005⫾0.003
`ND
`ND
`ⱕ0.010
`
`mM
`
`0.34⫾0.08
`0.36⫾0.05
`1.1⫾0.3
`12⫾2
`ND
`ND
`0.5⫾0.3
`
`(kcat/Km,␣KG,37°C)
`
`Efficiency
`
`(37°C)
`Km,NADPH
`
`(37°C)
`Km,␣KG
`
`1.59⫾0.09
`1.60⫾0.07
`4.2⫾0.3
`0.34⫾0.02
`ⱕ0.019
`ⱕ0.016
`0.019⫾0.001
`s⫺1
`
`kcat(37°C)
`
`cND,notdetermined.
`bFromRef.44.
`aFromRef.43.
`
`R132G
`R132C
`R132H
`R100Q
`A134D
`H133Q
`WT
`
`IDH1
`
`Valuesresultfromfitsofkineticdatausingtwodifferentenzymepreparations.Thestandarderrorisdeterminedfromthedeviancefromthesehyperbolicfits(Fig.4,supplementalFig.S4).
`Kineticparametersfortheneomorphicreaction,conversionof␣KGtoD2HG,catalyzedbyIDH1
`Table2
`
`48(Gly)
`86(Cys)
`118(His)
`148(Arg)
`148(Arg)
`148(Arg)
`148(Arg)
`
`0(Gly)
`49(Cys)
`8(His)
`⫺14(Arg)
`⫺14(Arg)
`⫺14(Arg)
`⫺14(Arg)
`
`atresidue132,Å3b
`volumeofsidechain
`
`vanderWaals
`
`hydrophobicityof
`
`residue132a
`
`Relative
`
`Rigel Exhibit 1051
`Page 7 of 13
`
`

`

`Catalytic efficiency of IDH1 mutants
`
`Figure 7. Comparisons of catalytic efficiency by IDH1 with mutations at residue 132. The observed rate constants (kobs) were calculated from the linear
`range of the slopes of plots of concentration versus time, and then fit to a hyperbolic equation to generate kcat and Km values. All experiments were performed
`at 37 °C. These catalytic parameters result from fits of kinetic data resulting from two different enzyme preparations to ensure reproducibility. A, relative
`catalytic efficiencies (kcat/Km) of the conversion of ICT to ␣KG using Km values for ICT are plotted against relative hydrophobicity (43). B, relative catalytic
`efficiencies (kcat/Km) of the conversion of ICT to ␣KG using Km values for ICT are plotted against van der Waals volume (44). C, relative catalytic efficiencies
`(kcat/Km) of the conversion of ␣KG to D2HG using Km values for ␣KG are plotted against relative hydrophobicity (43). D, relative catalytic efficiencies (kcat/Km) of
`the conversion of ␣KG to D2HG using Km values for ␣KG are plotted against van der Waals volume (44).
`
`Table 3
`Kinetic parameters for the normal reaction, conversion of ICT to ␣KG, catalyzed by IDH1
`Values result from fits of kinetic data using two different enzyme preparations. The standard error is determined from the deviance from these hyperbolic fits (supplemental
`Fig. S5).
`Km,ICT
`kcat
`(37 °C)
`(37 °C)
`sⴚ1
`mM
`1.21 ⫾ 0.08
`3.6 ⫾ 0.6
`10.4 ⫾ 0.2
`5.7 ⫾ 0.4
`9.2 ⫾ 0.3
`0.8 ⫾ 0.2
`7.2 ⫾ 0.4
`1.1 ⫾ 0.2
`0.047 ⫾ 0.001
`1.5 ⫾ 0.1
`
`Relative hydrophobicity
`of residue 132a
`
`van der Waals volume of side
`chain at residue 132, Å3 b
`
`97 (Trp)
`41 (Ala)
`⫺10 (Gln)
`⫺23 (Lys)
`⫺28 (Asn)
`
`163 (Trp)
`67 (Ala)
`114 (Gln)
`135 (Lys)
`96 (Asn)
`
`IDH1
`
`R132W
`R132A
`R132Q
`R132K
`R132N
`a From Ref. 43.
`b From Ref. 44.
`
`Efficiency
`(kcat/Km,ICT, 37 °C)
`mMⴚ1 sⴚ1
`0.34 ⫾ 0.06
`1.8 ⫾ 0.1
`12 ⫾ 3
`7 ⫾ 1
`0.031 ⫾ 0.008
`
`Table 4
`Kinetic parameters for the neomorphic reaction, conversion of ␣KG to D2HG, catalyzed by IDH1
`Values result from fits of kinetic data using two different enzyme preparations. The standard error is determined from the deviance from these hyperbolic fits (supplemental
`Fig. S6).
`Km,␣KG
`kcat
`(37 °C)
`(37 °C)
`sⴚ1
`mM
`0.54 ⫾ 0.01
`0.82 ⫾ 0.08
`0.37 ⫾ 0.01
`0.11 ⫾ 0.02
`4.7 ⫾ 0.2
`0.26 ⫾ 0.04
`0.57 ⫾ 0.02
`0.61 ⫾ 0.07
`0.79 ⫾ 0.06
`10 ⫾ 2
`
`Relative hydrophobicity
`of residue 132a
`
`van der Waals volume of side
`chain at residue 132, Å3 b
`
`97 (Trp)
`41 (Ala)
`⫺10 (Gln)
`⫺23 (Lys)
`⫺28 (Asn)
`
`163 (Trp)
`67 (Ala)
`114 (Gln)
`135 (Lys)
`96 (Asn)
`
`IDH1
`
`R132W
`R132A
`R132Q
`R132K
`R132N
`a From Ref. 43.
`b From Ref. 44.
`
`Efficiency
`(kcat/Km,␣KG, 37 °C)
`mMⴚ1 sⴚ1
`0.659 ⫾ 0.007
`3.4 ⫾ 0.6
`18 ⫾ 3
`0.9 ⫾ 0.1
`0.08 ⫾ 0.02
`
`tural comparisons for R100Q IDH1 and R140Q IDH2 will be
`important for characterizing any mechanistic differences
`between these homologous mutations. Currently, crystal struc-
`tures of R140Q IDH2 are limited to complexes with inhibitors
`(49). R100Q IDH1 has been characterized as a D2HG-producer,
`but our data suggests this mutant does so only weakly.
`Similarly, R132K IDH1 is homologous to R172K IDH2, the
`second most common IDH2 mutation identified in AML (50).
`Thus the relative catalytic inefficiency of D2HG production by
`R132K IDH1 (Table 4) was also surprising. Kinetic and struc-
`
`tural analysis of R132K IDH1 and R172K IDH2 will also be
`critical for understanding any functional differences between
`these homologous mutations. Nearly negligible in vitro D2HG
`production by R132K and R100Q IDH1 may explain why these
`mutations are rare (or not identified) in gliomas, despite being
`frequently observed as R172K and R140Q IDH2 in AML.
`As noted, Km values for D2HG production for some IDH1
`mutants are higher than physiologically relevant ␣KG concen-
`trations. However, IDH1 mutations are found heterozygously
`in tumors, and a caveat to this work in that mutant IDH1
`
`7978 J. Biol. Chem. (2017) 292(19) 7971–7983
`
`Rigel Exhibit 1051
`Page 8 of 13
`
`

`

`homodimers were studied. For IDH1 mutant/WT hetero-
`dimers, local concentration of ␣KG may be much higher
`due to production of this metabolite at the WT IDH1 mono-
`mer, particularly if substrate channeling occurs. Furthermore,
`Km values may be lower overall due to favorable substrate bind-
`ing at the WT monomer of the heterodimer. Ward et al. (28)
`have shown that D2HG production by IDH1 mutations in cells
`is increased if WT IDH1 activity is retained. However, in cases
`of very high Km values such as those observed for R100Q IDH1,
`it is possible that the reaction measured is not physiologically
`relevant, even in heterodimer

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