throbber
Article
`
`pubs.acs.org/JACS
`
`Protodeboronation of Heteroaromatic, Vinyl, and Cyclopropyl
`Boronic Acids: pH−Rate Profiles, Autocatalysis, and
`Disproportionation
`Paul A. Cox,† Andrew G. Leach,§ Andrew D. Campbell,‡ and Guy C. Lloyd-Jones*,†
`†School of Chemistry, University of Edinburgh, Joseph Black Building, David Brewster Road, Edinburgh EH9 3FJ, United Kingdom
`‡Pharmaceutical Technology and Development, AstraZeneca, Silk Road Business Park, Macclesfield SK10 2NA, United Kingdom
`§School of Pharmacy and Biomolecular Sciences, Liverpool John Moores University, Byrom Street, Liverpool L3 3AF, United
`Kingdom
`*S Supporting Information
`
`ABSTRACT: pH−rate profiles for aqueous−organic protodeboronation of 18
`boronic acids, many widely viewed as unstable, have been studied by NMR and
`DFT. Rates were pH-dependent, and varied substantially between the boronic
`acids, with rate maxima that varied over 6 orders of magnitude. A mechanistic
`model containing five general pathways (k1−k5) has been developed, and
`together with input of [B]tot, KW, Ka, and KaH, the protodeboronation kinetics
`can be correlated as a function of pH (1−13) for all 18 species. Cyclopropyl and
`vinyl boronic acids undergo very slow protodeboronation, as do 3- and 4-pyridyl
`boronic acids (t0.5 > 1 week, pH 12, 70 °C). In contrast, 2-pyridyl and 5-thiazolyl
`boronic acids undergo rapid protodeboronation (t0.5 ≈ 25−50 s, pH 7, 70 °C),
`via fragmentation of zwitterionic intermediates. Lewis acid additives (e.g., Cu, Zn
`salts) can attenuate (2-pyridyl) or accelerate (5-thiazolyl and 5-pyrazolyl)
`fragmentation. Two additional processes compete when the boronic acid and the
`boronate are present in sufficient proportions (pH = pKa ± 1.6): (i) self-/autocatalysis and (ii) sequential disproportionations of
`boronic acid to borinic acid and borane.
`
`■ INTRODUCTION
`Boronic acids are key reagents in synthesis,1 and ubiquitous in
`classic processes such as, inter alia, Suzuki−Miyaura,2 oxidative
`Heck,3 Chan−Evans−Lam,4 and Liebeskind−Srogl couplings,5
`and addition to enones,6 carbonyls, and imines.7 Boronic acid
`decomposition, notably by in situ protodeboronation,1
`compromises
`reaction efficiency, and motifs
`such as 2-
`heteroaryl,8 vinyl,9 and cyclopropyl10 are sometimes trouble-
`some. As a consequence, a range of techniques have been
`developed to mitigate decomposition during coupling:1,11 these
`include highly tuned catalysts,12 the use of additives (e.g., Cu,
`Zn, and Ag salts) sometimes proposed to act by trans-
`metalation,13 masked reagents,14 and slow release of
`the
`boronic acid in situ11 from MIDA boronates15 and trifluor-
`oborates.16
`Given the importance of boronic acids in Suzuki−Miyaura
`coupling, a process that is frequently conducted in aqueous−
`organic solvent media,16c general mechanistic understanding of
`direct aqueous protodeboronation is
`surprisingly limited.
`Moreover, nearly all studies have focused on substituted
`phenylboronic acids.1,12,17 The most detailed investigation was
`reported by Kuivila, who measured the protodeboronation
`kinetics of a series of ArB(OH)2 species (Ar = o,m,p-X-C6H4; X
`= MeO, Me, Cl, and F) in aqueous buffers at 90 °C, with initial
`ArB(OH)2 concentrations in the range 3−5 mM.17c−f By
`
`analysis of pH−rate profiles (between pH 1.0 and 6.7), two
`pathways were identified, Scheme 1. The first was a specific
`acid-catalyzed process (k1), shown to proceed via aromatic
`electrophilic substitution of B by H. The second pathway was
`found to be base-catalyzed, and proposed to proceed via
`hydrolysis (k2) of the boronate anion ([ArB(OH)3]−). This
`latter species is generated in a pH-determined equilibrium
`involving association of water (Ka)18 or hydroxide (Ka/KW)
`with the boronic acid. A key issue is that the kinetics were
`measured by UV−vis
`spectroscopy and could not be
`determined above pH 6.7 due to the onset of UV-dominating
`boronic acid oxidation processes.17f As a consequence of the
`pH being substantially below that required to effect significant
`conversion of
`the boronic acid to the boronate,
`the rate
`constant (k2) was calculated from kobs, using an estimated value
`for Ka. In addition to the elucidation of the two major pathways
`(k1 and k2) for aqueous protodeboronation, Kuivila’s studies
`also identified that electron withdrawing groups, at para or meta
`positions on the aromatic ring, attenuate protodeboronation
`rates, via both pathways. However, while these studies were
`extensive,17c−f they were conducted long before the ascendency
`the Suzuki−Miyaura reaction.2
`of
`In other words,
`the
`
`Received: March 30, 2016
`Published:
`June 29, 2016
`
`© 2016 American Chemical Society
`
`9145
`
`DOI: 10.1021/jacs.6b03283
`J. Am. Chem. Soc. 2016, 138, 9145−9157
`
`PETITIONER NPC EX. 1023
` Page 1 of 13
`
`

`

`Journal of the American Chemical Society
`
`Scheme 1. Kuivila Mechanisms (k1, k2) for ArB(OH)2
`Protodeboronation in Aqueous Acidic (k1)17c and Basic
`(k2)17f Mediaa
`
`involving boronate [ArB(OH)3]−
`aAlso shown is a third pathway,
`deprotonation and C−B protolysis (k3), proposed by Perrin17k for
`substrates with 2,6-disubstitution (X = F, Cl, Br, CF3), and an
`uncatalyzed pathway (k4) involving direct reaction with (autoionized)
`water.
`
`importance of detailed study of the base-catalyzed process,
`across the full alkaline pH range (i.e., well above pH 6.7), was
`not yet apparent.
`Indeed,
`it was only rather recently that the kinetics of
`protodeboronation of arylboronic acids have been studied
`under basic conditions. In 2002, Frohn measured protodeboro-
`nation rates of various polyfluorophenyl boronic acids in
`aqueous pyridine, and in aqueous basic methanol, concluding
`that the mechanism involved protolysis of either the boronate
`(i.e., k2) or the conjugate base of the boronic acid.17i In 2003
`Cammidge reported in detail on the effect of various anhydrous
`and aqueous−organic media in the protodeboronation of 2,3-
`difluoro-4-heptyl-6-tolyl boronic acid mediated by CsF,17j
`concluding that aqueous protolysis of
`the corresponding
`boronate was involved. In 2010 Buchwald used calorimetry to
`measure protodeboronation kinetics of a series of substituted
`2,6-difluorophenyl boronic acids (3-F, 3-OBu, 4-F, and 4-H) in
`a biphasic basic aqueous medium (aq K3PO4/THF).12 Perrin
`extended this study,
`including other electronegative 2,6-
`disubstituents, Cl, Br, and CF3,17k which led to the proposal
`of a new, i.e., non-Kuivila type, mechanism involving specific-
`base-mediated protolysis (k3) of the boronate anion ([ArB-
`(OH)3]−). The process was reported to only occur with
`boronic acids bearing a substituent at both ortho-positions (i.e.,
`2,6-disubstitution).17k
`Despite the core role of heteroaromatic boronic acids in
`synthesis and discovery, and the propensity for many to
`undergo protodeboronation, during storage17i and in cou-
`pling,8,12 there is a near-complete absence of the kinetic data
`requisite for their behavior to be compared and contrasted.
`Thus, while it is known empirically, or anecdotally, that certain
`heteroaromatic boronic acids are much more prone to
`protodeboronation than others,1,8,11b it is not clear whether
`overall they behave similarly to substituted phenylboronic acids,
`i.e., displaying the simple acid- and base-catalyzed pH
`relationships (k1, k2) identified by Kuivila, or whether there
`are more complex pH dependencies for some classes of
`heteroaromatic boronic acids, for example involving heterocycle
`basicity, or other pathways, such as the specific-base-mediated
`protolysis (k3) identified by Perrin.17k Indeed, it is not even
`clear for an individual class of heteroaromatic boronic acid
`
`Article
`
`whether extremes of pH (low or high) are to be avoided, or are
`beneficial, in terms of stability.
`For all of the above reasons, we set out to study the intrinsic
`aqueous protodeboronation of a range of heteroaromatic (2−
`17), vinyl (18), and cyclopropyl (19) boronic acids,
`in a
`homogeneous organic−aqueous medium. Herein, we report the
`overall kinetics of
`their protodeboronation, but more
`importantly also show how the resulting pH−rate profiles can
`be simulated and analyzed using a general kinetic model. The
`model extends beyond the basic Kuivila processes (k1, k2), by
`including the Perrin mechanism (k3), plus three additional
`protolysis processes (k2cat, k4, and k5, vide infra), the requisite
`pH-dependent speciation equilibria for boronic acid association
`with water (Ka), and, if required, the protonation state of basic
`heterocycles (KaH). The model can be used in two ways: (i) as
`a general exploratory tool, with manual input of the requisite
`pH, concentrations, rates, and equilibrium constants, or (ii) as a
`means to fit experimental data, through automated numerical
`iteration of rate and equilibrium constants (including Ka and
`KaH), provided that the rate data has been acquired over a
`suitably wide pH range. To assist in application of the model, a
`preconfigured spreadsheet is provided as part of the Supporting
`Information.
`the
`The model provides the basis for quantification of
`dominant protodeboronation processes occurring for different
`boronic acid species, at different pH and substrate concen-
`trations. Thus,
`the impact of pH on the kinetics allows
`identification of new mechanistic regimes, and these mecha-
`nisms can then be explored in more detail using kinetics,
`isotopic labeling, effects of additives, and computation. Overall
`the study has facilitated the following: (i) classification of the
`reactivity imparted by 16 different heterocyclic structures (2−
`17) between pH 1 and 13; (ii) elucidation and investigation of
`new protodeboronation mechanisms and a competing dis-
`proportionation process; (iii) preliminary details on origins of
`the (de)stabilizing effect of additives such as Zn and Cu salts on
`some heterocyclic boronic acids; and (iv) identification of
`substrate-specific pH-stability zones, in which even notoriously
`unstable boronic acids, e.g., 2-pyridyl, can be stable for a few
`hours at 70 °C. This information will aid a more informed
`the preparation, storage,17i and
`choice of conditions for
`application of boronic acids in synthesis,1 as well as a means
`to induce their deliberate19 protodeboronation.
`■ RESULTS AND DISCUSSION
`1. Protodeboronation via Kuivila Mechanisms (k1 and
`k2). In preliminary studies we confirmed that the protodeboro-
`nation of a simple para-substituted phenyl boronic acid could
`be satisfactorily analyzed in situ by 1H/11B NMR under aqueous
`conditions. With p-anisyl boronic acid (1),17c,f protodeborona-
`tion kinetics were determined in water at 90 °C, without a
`malonate buffer.17c,f The aqueous association constant (pKa 1 =
`9.10; 90 °C) was determined by 11B NMR pH titration.
`Control experiments confirmed that basic solutions of 1
`became pale brown in color, with large increases observed in
`UV−vis absorption spectra, as reported by Kuivila.17f However,
`the NMR spectra of such samples were unaffected: there was
`no sign of mequinol (p-hydroxy anisole),
`the anticipated
`product of oxidation of 1, or indeed anything other than the
`time-average signal from [1/1OH] and the protodeboronation
`products, anisole and boric acid. The trace quantities of
`oxidative side product(s) are thus intensely UV-active, and
`possibly polymeric.
`
`9146
`
`DOI: 10.1021/jacs.6b03283
`J. Am. Chem. Soc. 2016, 138, 9145−9157
`
`PETITIONER NPC EX. 1023
` Page 2 of 13
`
`

`

`Journal of the American Chemical Society
`
`Using HCl and KOH to explore the acid (pH 1−3) and base
`(pH 11−13) regimes, protodeboronation kinetics were
`analyzed by nonlinear regression of the exponential decays
`observed for [1/1OH]. The second-order rate constants (k1 and
`k2) are given in Table 1, entry 1. Kuivila’s value for the limiting
`
`Table 1. Protodeboronation 1 and 2 at 90 °C
`
`k2 M−1 s−1
`k1 M−1 s−1
`a
`pKa
`ArB(OH)2
`entry
`3.9 × 10−8b
`0.68 × 10−4b
`9.10
`1
`1
`8.4 × 10−6f
`1.1 × 10−4e
`9.60d
`2c
`1
`1.4 × 10−8b
`3.3 × 10−6b
`8.91
`2
`3
`a11B NMR pH titration. b11B NMR at pH 1−3 and 11−13. cFrom refs
`17c and f. dEstimated in ref 17f. eExtrapolated from data in ref 17c.
`fFrom kobs at pH 6.7 (25 °C, uncorrected) and estimated Ka.
`
`rate constant at high pH (k2, entry 2) was obtained by pH−rate
`extrapolation and is 2 orders of magnitude too large.17k This
`arises from the conflation of an overestimated pKa for 1 (9.60),
`with an uncorrected pH (25 °C)20 for the kobs determination at
`90 °C, and reinforces the value of full pH range rate profiling.
`Moving to the protodeboronation of heterocycles, 3-
`thienylboronic acid (2) was chosen for initial studies, on the
`basis of its solubility, relative stability, and low basicity. Second-
`order rate constants (k1 and k2; Table 1, entry 3) were
`determined under the same conditions (50 mM, H2O, 90 °C)
`as for p-anisyl boronic acid 1. Within the limits of the pH range
`explored (pH 1−13), there was no detectable contribution by
`the base-catalyzed boronate mechanism (k3), or direct reaction
`of the boronic acid with H2O (k4; Scheme 1),21 although both
`mechanisms (k3 and k4) were found to be important with some
`heterocycles, vide infra.
`3-Thienyl boronic acid 2 is less susceptible to aromatic
`electrophilic substitution (k1)22 than 1, but the boronates (1OH
`and 2OH) are of similar reactivity (k2, Table 1, entries 1 and 3).
`Computational studies on this process identified rate-limiting
`C-protonolysis of the boronate (2OH, Figure 1, upper structure)
`by water. As found for all of the boronates studied, the three
`hydroxyl groups in 2OH preferentially adopt a distinctive
`triskelion conformation. Hydrogen bonding of the boronate by
`incoming water results in the C−B bond stretching such that, in
`the transition state (TS), the 3-thienyl carbanion is midpoint in
`its translation from the Lewis acid (B(OH)3) to the Brønsted
`acid (OH).
`2. Autocatalyzed Protodeboronation (k2cat). If proto-
`deboronation of 2 arises solely by the Kuivila mechanisms (k1,
`k′2), (k′2 = k2[H2O]) the empirical rate equation predicts a
`simple pH−log kobs rate profile, Figure 2. Specifically, for k′2
`(see solid line, Figure 2), the rate should rise and then reach a
`plateau as the pH extends above the pKa of the boronic acid 2
`(pKa = 8.91 at 90 °C). The data deviates from this theoretical
`curve, with the deviation being greatest at pH 8.9. This
`deviation was found for many of the boronic acids studied, vide
`infra, and indicates that there is an additional protodeborona-
`tion process that augments k′2, but with a distinctly different
`pH profile. A pronounced concentration dependence was
`
`Article
`
`Figure 1. DFT (M06L/6-311++G**)23−27 transition state structures
`for protonolysis of 3-thienylboronate 2OH by water (k2, top) and by
`boronic acid (2, k2cat, bottom) to generate 2H and [B(OH)4]−.
`Distances are shown in Å and free energies are given in kcal/mol.
`
`Figure 2. Effect of pH and concentration on the rate of
`protodeboronation of thienyl boronic acid 2. Dashed lines: kobs =
`((k′2 + k2cat[B]tot)/(1 + 10(pKa−pH))) + ((k1/10pH)/(1 + 10(pH−pKa)))
`values as Table 1, entry 3, plus k2cat = 6.2 × 10−5 M−1 s−1; for solid
`line, k2cat = 0. Inset shows rate profile in D2O (red), using an x-axis
`scale of pD + ΔpKW to account for the change in water autoionization
`(KW).30 pKa 2 (11B NMR pH titration, 90 °C): 8.91 (H2O) and 9.68
`(D2O).
`
`noted, with the deviation from k′2 becoming greater as the
`boronic acid concentration is raised (compare 0.05 to 0.40 M,
`Figure 2). As the maximum deviation occurs when pH = pKa
`(8.91), where the proportions of [RB(OH)3]− and RB(OH)2
`are identical, analogous to that of a bimolecular Job-plot,28 this
`suggests self-catalysis (k2cat). To account for the approximately
`pseudo-first-order kinetics (vide infra), the product (B(OH)3)
`needs to be a similarly effective autocatalyst,29 as the boronic
`acid is a self-catalyst.17l This was confirmed by addition of 350
`mM 10B(OH)3 to the protodeboronation of 50 mM 2. At pH
`
`9147
`
`DOI: 10.1021/jacs.6b03283
`J. Am. Chem. Soc. 2016, 138, 9145−9157
`
`PETITIONER NPC EX. 1023
` Page 3 of 13
`
`

`

`Journal of the American Chemical Society
`
`8.90, 11B NMR analysis afforded the same pseudo-first-order
`rate constant as for 400 mM 2 alone ([B]tot = 400 mM in both
`cases). Using an augmented rate equation containing k2cat
`allows satisfactory kinetic simulations (dashed lines in Figure
`2). DFT transition state energies confirmed that water can be
`replaced by boronic acid 2 (Figure 1, lower structure) or boric
`acid as proton sources, and KIEs determined from reactions in
`D2O (Figure 2, inset) confirm rate-limiting proton-transfer.
`3. B−OH-Catalyzed Disproportionation. After some
`preliminary investigation with heterocyclic boronic acids, a
`mixture of H2O/1,4-dioxane (1:1) at 70 °C was found to
`provide the best combination of substrate solubility (50 mM)
`while remaining fully homogeneous at the wide range of base,
`salt, buffer, additive, and acid concentrations required to
`usefully explore the pH range (1−13). With higher initial
`concentrations of boronic acid, some substrates displayed small
`deviations from pseudo-first-order protodeboronation kinetics
`when carefully monitored at pH = pKa ± 0.1.
`This deviation had already been observed with thienyl
`boronic acid 2 in H2O at 90 °C, vide supra, and together with
`tests for boryl exchange at the aryl ring, Scheme 2, indicated
`that processes in addition to k2cat occur when both [RB-
`(OH)3]− and RB(OH)2 are present at high concentration (pH
`= pKa).
`
`Scheme 2. 10B/11B Exchange: Reaction of 11B(OD)3 (0.2 M)
`with [10B]-2OD (0.2 M) To Generate Small Quantities of
`[11B]-2/[11B]-2OD and 10B(OD)3 (10/11B NMR, D2O, 90 °C)a
`
`aRegioisomer 3 protodeboronates without significant disproportiona-
`tion.
`
`11B NMR analyses of this process indicated other transient
`minor (≤5%) species
`to be present, and the pH-shift
`dependency of one of these species suggested it to be a
`borinate,31 [(3-thienyl)2B(OH)2]−, generated by disproportio-
`nation of 2/2OH. Regioisomeric 2-thienyl boronic acid (3) was
`found to undergo much faster protodeboronation (k2 and
`k2cat) than 3-thienyl 2, with no significant disproportionation.
`In contrast, 2-furyl boronic acid (4, 1:1 d8-dioxane/D2O, 0.5
`equiv of KOD; pD = pKa) underwent disproportionation faster
`than protodeboronation when the concentration was raised to
`0.4 M, Figure 3.
`The temporal evolution (1H NMR) indicated a two stage
`process, initially giving difurylborinic acid (R2B(OD)), which
`disproportionates further to give the trifurylborane (R3B, Figure
`3). Over a period of days, all species protodeboronate, directly
`or indirectly, to give boric acid and furan (4D).33 DFT studies
`investigated a number of mechanisms for the aryl transfer
`between the boron centers. These included an analogue of the
`autocatalysis mechanism, Figure 1, TS 2OH (k2cat), in which the
`furyl anion transfers to the Lewis acid B, rather than Brønsted
`
`Article
`
`Figure 3. Upper: temporal concentration data for disproportionation
`of 2-furyl boronic acid 4 (1:1 d8-dioxane/D2O, 0.5 equiv of KOD, 70
`°C). Circles: data. Solid lines: simulation (see SI for details). 2H1-furan
`(RD = 4D) is volatile under the reaction conditions and was not
`monitored. Lower: transition states (DFT) for disproportionation to
`difurylborinic acid and trifurylborane.32
`
`acid H, in an H-bonded [4 + 4OH] intermediate. However, the
`computed barrier for transfer to B (22.5 kcal/mol) is greater
`than that for H (20.3 kcal/mol), and protodeboronation would
`dominate if solely these isomeric transition states are operative.
`An alternative process was therefore considered in which a
`cyclic boroxine-ate complex (Figure 3, X = furyl or OH, upper
`structure) facilitates aryl-migration across the ring. The rate-
`limiting barriers for aryl migration are computed to be 19.1−
`19.4 kcal/mol for 2-furyl 4OH and 19.5−19.9 kcal/mol for 3-
`thienyl 2OH.
`Accurately computing free energies in solution for very
`different processes is challenging; nonetheless, the calculations
`suggest that migration dominates over protodeboronation for
`2-furyl 4, while the opposite is the case for 3-thienyl 2. The
`transition state structures suggest an electronic and steric
`component to this preference. Mulliken charges indicate the
`nonmigrating 2-furyl bears a larger negative charge than the
`equivalent group for 3-thienyl. Further,
`the preferred
`conformation of
`the transition state places the migrating
`group in close proximity to the other aromatic ring; larger rings
`or those with C−H groups adjacent to the boronic acid group
`will inevitably involve steric clashes, even more so when X =
`aryl. The subsequent step leading to triarylborane requires
`reaction of a borinic with a boronic acid and thus cannot
`involve a cyclic boroxine. Instead it is computed to proceed via
`a transition state involving a dimer (Figure 3, lower structure).
`The free energy barrier (+18.9 kcal/mol) is comparable to that
`for the first step, and thus consistent with the observed
`
`9148
`
`DOI: 10.1021/jacs.6b03283
`J. Am. Chem. Soc. 2016, 138, 9145−9157
`
`PETITIONER NPC EX. 1023
` Page 4 of 13
`
`

`

`Journal of the American Chemical Society
`
`Article
`
`Figure 4. pH−rate profiles (I−IV) for pseudo-first-order protodeboronation (kobs/s−1) of boronic acids 2−19 in 1:1 H2O/1,4-dioxane at 70 °C.
`Reactions analyzed in situ by 11B NMR were conducted in quartz NMR tubes to avoid [B(OH)4]− release from borosilicate NMR tubes. Circles:
`experimental data. Solid lines: simulation using model (Figure 5) with data from Table 2. Rates below log kobs = −7 are not modeled.
`
`Figure 5. General model for heterocycle protodeboronation (Z = basic nitrogen), individual pH−log kobs profiles, overall rate equation based on the
`three-state speciation [neutral (X), N-protonated (XH+), and boronate (XOH)], and example combined pH−log kobs profiles (i, ii, iii).
`
`transient accumulation of diarylborinic acid. Using these
`models for
`reversible furyl
`transfer, with an irreversible
`
`9149
`
`DOI: 10.1021/jacs.6b03283
`J. Am. Chem. Soc. 2016, 138, 9145−9157
`
`PETITIONER NPC EX. 1023
` Page 5 of 13
`
`

`

`Journal of the American Chemical Society
`
`Article
`
`Table 2. Equilibrium and Rate Constants Employed in the pH Simulation (See Overall Rate Equation in Figure 5) for
`Protodeboronation of Heteroaromatic, Vinyl, and Cyclopropyl Boronic Acids 2−19 in 1:1 H2O/1,4-Dioxane at 70 °C
`
`a
`pKaH
`
`b
`
`e
`
`log k′4
`
`log k′5
`
`log k′3
`log k′2
`c
`log k2cat
`log k1
`pKa
`RB(OH)2
`entry
`≤−5.32d
`−4.21
`−6.01
`−5.95
`11.04
`2
`1
`≤−3.07d
`≤−3.54d
`−3.92
`−4.44
`10.38
`3
`2
`≤−4.12d
`−3.32
`−4.98
`−3.71
`10.29
`4
`3
`≤−1.93d
`≤−2.07d
`−2.34
`−3.41
`9.64c
`5
`4
`−3.72
`−3.27
`≤−6.02d
`−3.16
`11.61
`6
`5
`−2.62
`−2.00
`−4.01
`−5.51
`10.45
`7
`6
`≤−4.75d
`−3.44
`−5.04
`≤−4.82d
`8.18
`8
`7
`≤−5.59d
`−3.60
`−6.05
`≤−3.10d
`9.76
`9
`8
`≤−5.78d
`−3.78
`−5.93
`≤−3.76d
`8.94
`10
`9
`≤−6.60d
`−3.95
`−6.24
`≤−5.64d
`9.90
`11
`10
`≤−2.21d
`≤−1.58d
`−2.02
`≤−2.60d
`8.41c
`12
`11
`≤−3.94d
`−3.49
`−3.87
`≤−4.22d
`8.82
`13
`12
`≤−3.93d
`−3.19
`−4.35
`≤−4.58d
`9.11
`14
`13
`≤−3.74d
`≤−0.67d
`≤−4.26d
`≤−0.74d
`10.76c
`15
`14
`≤−6.04d
`≤−2.33d
`≤−6.72d
`≤−1.36
`9.54c
`16
`15
`≤−4.82d
`≤−1.84d
`≤−6.32d
`≤−2.06d
`8.49c
`17
`16
`≤−5.73d
`−3.90
`−6.02
`−5.57
`11.21
`18
`17
`≤−5.91d
`−5.01
`−6.01
`−5.95
`12.72
`19
`18
`apKaH in parentheses determined by 1H NMR pH titration at 25 °C; pKaH at 70 °C from iterative fitting of rate data (unless stated). bpKa by 11B
`NMR pH titration at 70 °C (unless stated). cpKa from iterative fitting of rate data. dThis value is not required for satisfactory simulation; greater
`values induce ≥5% change in the sum of square errors between simulation and data across the overall pH profile. epKaH fixed at empirical offset
`(−0.46 units) based on entries 5, 8, 11−16. fDue to the pKaH of 17 being <−1.06, this value is a substantial overestimate.
`
`1.26 (1.70)
`0.04e (<0.50)
`0.14e (<0.60)
`3.60 (4.22)
`3.36e (3.82)
`2.95e (3.41)
`1.62 (1.85)
`1.08 (1.62)
`1.36 (2.00)
`3.52 (3.86)
`3.16 (3.60)
`−1.06e (<−0.6)
`
`−6.48
`−6.60
`−6.38
`−5.44
`−5.71
`−5.71
`−6.89
`−1.83
`−3.78
`−4.28
`−1.60
`−2.41
`−2.75
`
`≤−5.68d
`≤−4.61d
`≤−4.83d
`≤−6.70d
`−5.77
`−7.07
`≤−4.22d
`≤−5.31d
`≤−5.84d
`≤−4.24d
`≤−4.52d
`≤−0.06d
`
`protodeboronation, predominantly via the boronic acid, a good
`fit to the experimental data was achieved by kinetic simulation
`(solid lines, Figure 3).
`4. General Model for Protodeboronation Kinetics.
`With suitable conditions established for analysis (50 mM
`substrate, 1:1 H2O/1,4-dioxane, 70 °C) the pseudo-first-order
`reactions of boronic acids 2−19 were
`kinetics (kobs) of
`measured as a function of pH (1−13), Figure 4, profiles I−
`IV, see SI for full details. Rates were pH-dependent, and varied
`substantially between the boronic acids, with rate maxima that
`varied over 6 orders of magnitude (half-lives ranging from
`seconds to weeks). For some substrates, specific pH ranges
`reduced the protodeboronation rates to such an extent that, for
`reasons of accuracy, they were omitted from the log(kobs)−pH
`analyses. An arbitrary threshold of log kobs ≥ 7.0 (half-life ≤ 80
`days) was set for data inclusion (see dashed lines in profiles I−
`IV, Figure 4).
`To analyze the pH-dependency of the protodeboronation
`reactions of heterocyclic boronic acids, we developed a general
`model, Figure 5. Aiming to keep as minimal a model as
`required, we began with the Kuivila processes, and added
`further processes, when necessary, as
`the analysis of
`heterocyclic boronic acids 2−17 evolved. To simplify the
`discussion, the overarching model is presented in advance of
`the analysis of the protodeboronation characteristics of 2−17.
`As many of the heterocyclic systems studied are basic, the
`model
`includes, when appropriate,
`the pKaH of
`the N-
`
`in addition to the pKa
`
`protonated form of the heterocycle,
`for aqueous association at boron.
`The kinetics are determined by a 3-fold speciation of the
`boronic acid (X): as an N-protonated form (XH+; only for 6−
`17), a neutral form (X), and a boronate form (XOH). Specific
`protodeboronation processes occur from the three speciation
`states. For the neutral form (X), there is the acid-catalyzed
`Kuivila process (k1), and a pH-independent direct reaction with
`water (k4). For the latter, although this is not kinetically
`differentiated in the model, basic heterocyclic boronic acids can
`engage in a pre-equilbrium (K4) with autoionized water to
`generate a zwitterionic adduct (XZW). For the boronate form
`(XOH),
`there are the base-catalyzed Kuivila process (k2),
`concentration-dependent autocatalysis (k2cat) occurring with
`rate maximum when pH = pKa, and the Perrin mechanism
`involving base-catalyzed protonolyis (k3). For
`the latter,
`although this is not kinetically differentiated in the model, the
`boronate can engage with hydroxide in a pre-equilbrium (K3)
`to generate the dianion (XO‑). Finally, for the N-protonated
`form (XH+), this being distinct from the zwitterion (XZW) due
`to the presence of a boronic acid, not a boronate, there is a
`direct protodeboronation by water (k5).
`As
`indicated graphically beside the protodeboronation
`mechanisms in the model (Figure 5), each of the six processes
`(k1, k2, k2cat, k3, k4, k5) has a distinct pH−log kobs profile. It can
`be instructive to consider hypothetical combinations of selected
`pH−log kobs
`relationships; examples (i,
`ii, and iii) are
`
`9150
`
`DOI: 10.1021/jacs.6b03283
`J. Am. Chem. Soc. 2016, 138, 9145−9157
`
`PETITIONER NPC EX. 1023
` Page 6 of 13
`
`

`

`Journal of the American Chemical Society
`
`preconfigured in the spreadsheet provided in the SI. By
`mathematical combination of all six steps (see SI for full
`derivation) and calculation of the three-state speciation, an
`“overall rate equation” (Figure 5, center) can be generated. The
`equation allows analysis of
`the empirical rate (kobs) as a
`function of pH and boronic acid concentration, using up to
`nine constants: pKa, pKaH, k1, k′2, k′2cat, k′3, k′4, k′5, pKw, in
`which processes indicated by the term k′ contain amalgamated
`constants (e.g., K, k, and [H2O]). For nonbasic boronic acids,
`pKaH is nominally set to −5 to preclude XH+ speciation.
`Values for pKa and pKaH can be determined independently,
`or via the pH−log kobs simulation. Initial pKaH values for 6−17
`were determined by 1H NMR pH titration (1:1 H2O/1,4-
`dioxane, 25 °C), Table 2. An empirical correction for
`temperature (ΔpKaH 25−70 °C = −0.46) was determined
`from the pH−rate profile simulation. Generally, the B(OH)2
`unit slightly decreases the basicity relative to the parent
`heterocycle.34 For example, 2-, 3-, and 4- pyridyl boronic acids
`(15, 9, 10) have pKaH values (25 °C) of 3.86, 4.22, and 3.82,
`compared to that of 4.38 for pyridine, see SI. The pKa values for
`most of the boronic acids were measured by 11B NMR pH
`titration (1:1 H2O/1,4-dioxane, 70 °C); the pKa of the more
`reactive species (5, 12, 15−17) were estimated via the pH−log
`kobs simulation.
`5. pH−Rate Profiles Analysis of Boronic Acids 2−19.
`Using the rate equation in Figure 5, the pH−rate profiles for
`2−19 were simulated by automated iteration of rate and
`equilibrium constants, minimizing the sum square error (SSE)
`between predicted and observed data across the full profile (pH
`1−13). The SSE-minimized fitting constants are provided in
`Table 2. In all cases (2−19), only a subset of the six pathways
`(k1, k′2, k2cat, k′3, k′4, k′5) were required for satisfactory
`simulation (solid lines through data, Figure 4). The constants
`that are not required for simulation, but are in principle feasible,
`are reported as threshold values (≤) that induce a ≤5% change
`in the SSE.35
`Nonbasic Heterocycles (2−5). The thienyl (2, 3) and furyl
`(4) boronic acids (Figure 4, profile I) required only the Kuivila
`processes (k1, k′2)17c,k and autocatalysis (k2cat) for simulation.
`The 2-pyrrole boronic acid 5 required an additional pH-
`independent process (k′4)21 to fit the data between pH 3 and 6.
`This process is slow enough (k′4, half-life > 24 days) to be
`consistent with a water autoionization mechanism.36 Higher
`reactivity of 2- versus 3-thienyl and furyl boronic acids has been
`noted before, but only for acid-catalysis (k1).22 In all cases (2−
`5), the base-catalyzed process (k′2) is more efficient than the
`acid (k1), and the rates rise substantially through the series;
`above pH 11, 2-pyrrolyl 5 has a half-life of less than 3 min.
`Basic Heterocycles (6−17). The rate data obtained for 2-
`pyridyl boronic acids 15−17 (profile II) show a near inverse
`pH−rate profile compared to that of the nonbasic heterocycles
`2−5 (profile I) and required just a single term (k′4) for
`simulation. Maximum rates are attained when speciation
`disfavors the boronate (15−17OH; high pH) and pyridinium
`(15−17H+; low pH) forms. Thus, for the least basic 17 (pKaH
`<−0.6), the protodeboronation is not detectably attenuated by
`acid, even at pH 1. The much less reactive 5-pyrimidyl (8) and
`3-pyridyl (9) systems required base-catalysis (k′2 + k2cat) in
`addition to the neutral mechanism (k′4) for simulation. The 5-
`pyrazolyl and 5-thiazolyl boronic acids (13−15) were highly
`reactive, requiring both k′4 (neutral) and k′2 (basic) pathways
`for satisfactory simulation, with rates attenuated at pH below
`their pKaH (profile III). The 4-pyridyls (10, 11) are of similar
`
`Article
`
`they still
`in that
`reactivity to 3-pyridyl (9), but differ
`protodeboronate effectively at a pH substantially below their
`pKaH (profile III). This effect was satisfactorily simulated by
`including a direct (H2O-mediated) protodeboronation of the
`conjugate acid (k′5); the 4-pyridyl systems (10, 11) were the
`only species requiring this. The two remaining basic hetero-
`cycles, 4-pyrazolyl (6) and 4-isoxazolyl (7), gave the most
`complex profiles, requiring acid (k1), base (k′2/k2cat), base-
`catalyzed boronate (k′3), and neutral (k′4) pathways (profile
`IV; Figure 4). These were the only examples requiring the
`Perrin mechanism (k′3), a process that in Ar−B(OH)2 systems
`is proposed to require 2,6-disubstitution.17k
`Vinyl and Cyclopropyl Boronic Acids (18−19). Strongly
`acidic or basic solutions were required to effect any significant
`protodeboronation of the vinyl (18) and cyclopropyl (19)
`boronic acids (profile IV). Even at pH ≥ 11, the half-lives are
`weeks. A small selection of simple alkyl boronic acids (Me, c-
`Bu, and c-Hex) were also investigated. The extremely slow
`reactions (half-lives of months, see SI) made it difficult to
`clarify, by 1H/13C/11B NMR analysis, if protodeboronation was
`the major pathway of decomposition, rather than, for example,
`oxidation.
`6. Protodeboronation Mechanisms for Basic Hetero-
`cycles. To aid rationalization of the diverse range of pH
`for basic heterocycles 6−17, protodeboronation
`profiles
`mechanisms were explored by 11B NMR, DFT calculations
`(M06L/6-311++G**;23 see SI for details), and by testing the
`effect of additives.
`Zwitterionic Water Adducts (XZW). 2-Pyridyl boronic acids
`are notorious
`for
`their
`susceptibility to protodeborona-
`tion,1,8,11−13 as is found for 15−17. However, only a single
`neutral process (k′4) is required for simulation of the full pH−
`rate profile, indicating that the classic Kuivila-type acid- and
`base-catalyzed mechanisms (k1 and k′2) are negligible
`processes. Indeed H+/OH− act as powerful protodeboronati

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