Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Nov 7.
Published in final edited form as: Chem Commun (Camb). 2020 Oct 5;56(86):13105–13108. doi: 10.1039/d0cc05121g

Mobility of Lewis Acids within the Secondary Coordination Sphere: Toward a Model for Cooperative Substrate Binding

John J Kiernicki a, Emily E Norwine a, Myles A Lovasz a, Matthias Zeller b, Nathaniel K Szymczak a
PMCID: PMC7606458  NIHMSID: NIHMS1636055  PMID: 33016291

Abstract

Distance dependence of appended Lewis acids in N2H4 binding and deprotonation was evaluated within a series of zinc complexes. Variation of spacer-length to a tethered trialkylborane Lewis acid revealed distinct preferences for binding and stabilization of the resulting deprotonated N2H3 unit.

Graphical Abstract:

graphic file with name nihms-1636055-f0001.jpg


Acidic/basic residues within the secondary coordination sphere of metalloenzyme active sites are often critical for structure regulation and/or reactive-intermediate stabilization.1 These interactions occur in a dynamic protein environment where acidic residues are highly flexible and mobile. For example, in the active site of type II β-carbonic anhydrase, an aspartate residue (Asp44) gates reactivity to a Zn-OH: initially binding to Zn, then migrating 1.5 Å to hydrogen bond with Zn-OH2.2 Similarly, a recent report of Mo-nitrogenase revealed a dynamic multi-metallic cofactor, where the nitrogen reduction sequence is proposed to involve dynamic rotation and substrate/H-bonding interactions.3

Synthetic models containing pre-arranged secondary sphere groups can provide insight into the roles through which acidic groups facilitate substrate binding.4 While such models often use rigid molecular scaffolds to provide critical snapshots of donor/acceptor adducts, they do not capture mobility-dependent reactivity. Modelling mobility of an acidic residue within the secondary coordination sphere is synthetically challenging.5

Our group is working to evaluate how the precise structural, electronic, and cooperative modes in the secondary coordination sphere can be used to regulate reactivity.6 Recently, our lab investigated the role of ligand-appended acidic groups in homolytic bond scission of hydrazine by transition metal complexes.7 A key design aspect in those studies was the flexibility of the tethered trialkylborane Lewis acid, which enabled acid/base interactions to occur in both the primary (cooperatively with the transition metal) and secondary (independent of metal) spheres. A first-generation ligand, 3-BBNNNtBu, containing a 3-carbon alkyl tether between the Lewis acid and the bidentate ligand was synthesized by hydroboration of 2-(1-allyl-5-(tert-butyl)-1H-pyrazol-3-yl)-6-methylpyridine with 9-borabicyclo[3.3.1]nonane (BBN).8 While this tether length stabilized monoatomic ligands (e.g. -NH2) cooperatively with iron (Fig. 1), we obtained divergent results when attempting to sequester a diatomic substrate, cyanide.9 These disparate results highlight a Lewis acid dependence on substrate binding and illustrate a need to establish parameters (e.g. tether length) that maximize binding for a given substrate. Minimizing the energy requirement for cooperative acid-substrate-metal interaction will provide design cues that will enable the use of less-acidic Lewis acids, and ultimately facilitate product release—a challenge for catalytic turnover.

Fig. 1.

Fig. 1

Left: Conceptual design. Right: specific substrates investigated in this study and their potential binding modes.

To probe distance relationships between the Lewis acid and a given substrate, we evaluated a set of compounds where: 1) the Lewis acid is held constant (9-BBN), 2) the alkyl tether length is systematically varied from 2 to 4 methylene (-CH2-) units, and 3) the substrate contained variable bonding modes (Fig. 1). Zn was selected to ensure a consistent coordination geometry, and hydrazine (N2H4) was selected as the substrate (Fig. 1). Cooperative binding of N2H4 between the two Lewis acids, Zn2+ and trialkylborane, must involve coordinating each of the two lone pairs (i.e. μ−1,2-N2H4).

The two- and four-carbon length precursor ligands, vinylNNtBu and butenylNNtBu, were prepared by adapted literature procedures (see SI). Initial metalation strategies of the new ligands mirrored our synthesis of (3-BBNNNtBu)ZnBr2 (A-3; 3 denotes tether length, A denotes compound series).8 Stirring a CH2Cl2 solution of vinylNNtBu or butenylNNtBu with ZnBr2 furnished (vinylNNtBu)ZnBr2 and (butenylNNtBu)ZnBr2 as white powders (Fig. 2). Whereas late-stage hydroboration of (butenylNNtBu)ZnBr2 with 9-BBN proceeded (RT, THF, 18 hr) to afford (4-BBNNNtBu)ZnBr2 (A-4), (vinylNNtBu)ZnBr2 was obstinate to hydroboration. The molecular structures of (vinylNNtBu)ZnBr2 and (butenylNNtBu)ZnBr2 (Fig. 2B) determined by single crystal X-ray diffraction (SC-XRD) revealed a potential origin of the difference in hydroboration reactivity. Although hydroboration of vinyl substituents is generally facile with 9-BBN,10 (vinylNNtBu)ZnBr2 possesses two large steric moieties, the -C(CH3)3 and ZnBr2, that render the vinyl group inaccessible.11 To overcome this challenge, we pursued an early-stage hydroboration. Treating vinylNNtBu with 9-BBN generated 2-BBNNNtBu in situ, which was metalated with ZnBr2 in one-pot to afford (2-BBNNNtBu)ZnBr2 (A-2). Spectroscopically and structurally, A-2 and A-4 are similar to A-3. The distance between the two acidic centres, Zn and the trialkylborane, increases by ca. 1 Å for each additional –CH2– unit added to the tether length (Zn-B: A-2 = 4.76ave; A-3 = 5.82; A-4 = 6.57 Å).12 This trend suggests the system is well-suited for a distance-dependent cooperativity study.

Fig. 2.

Fig. 2

A) Synthesis of complexes A. B) Molecular structures (50% probability ellipsoids) of A-2 and A-4 as well as their vinyl and butenyl precursors, respectively. H-atoms not attached to alkenyl moieties are omitted for clarity.

Our previous studies revealed the three-carbon tether was ideally suited for cooperative binding of a wide range of μ−1,1 substrates;89 however, we hypothesized a distinct tether length would be needed for a μ−1,2 substrate. We used hydrazine (N2H4) to investigate this hypothesis. Previously, we demonstrated that the trialkylborane of A-3 could capture a single equivalent to N2H4 to afford (3-BBNNNtBu)ZnBr2(N2H4) (B-3), where the terminal -NH2 lone-pair is uncoordinated.13 The series of complexes B were synthesized by standard protocols and all share similar spectroscopic properties.8 Structurally, the coordination environment at Zn is unperturbed and the distances between the two acidic sites, Zn and boron, are variable and range from 4.80 (B-2) to 7.38 Å (B-4). All display weak intra- (B-2 and B-3) or intermolecular (B-4) NH…Br hydrogen bonding interactions14 that result in the N2H4 moieties being nearly equidistant (ave = 4.36 +/− 0.17 Å) to a Zn atom of the same, or an adjacent molecule.

Halide abstraction from complexes B with Tl+ afforded cationic complexes, [(n-BBNNNtBu)ZnBr(N2H4)][X], (C-2-C-4; X = OTf, PF6) and were subjected to SC-XRD studies (Fig. 3). Each complex displays a C1 symmetric tetrahedral bromido-Zn (τ4 = 0.83–0.85) chelated by the n-BBNNNtBu ligand. The fourth coordination site is occupied by a hydrazine ligand bridging to the appended trialkylborane (i.e. Zn-NH2NH2-BR3). Across the series of compounds, the Zn-N2H4 bond distance elongates from 2.0256(11) – 2.0892(15) Å (C-2 < C-3 < C-4) as the tether length to the trialkylborane increases (Δ = 0.055 Å). This trend is also observed, though to a lesser degree, in the B-N2H4 distance where C-2 is shorter (1.6463(18) Å) in comparison to C-3 and C-4 (1.675(2) and 1.668(2) Å, respectively). Both the R3B-N2H415 and Zn-N2H4 distances16 are comparable to related species. The interactions of N2H4 with the two Lewis acids, Zn and boron, force a nearly fixed distance from one another with variation of only 0.33 Å across the C series (contrasting with the B series; ΔZn-B = 2.58 Å), highlighting the accordion-like flexibility of the acidic trialkylborane.

Fig.3.

Fig.3

A) Reversible formation of C from B. B) Molecular structures (50% probability ellipsoids, only H-atoms attached to heteroatoms are displayed). B-3 is previously reported.8

The solid-state data of C-2 revealed both the shortest Zn-N and B-N contacts suggesting that this binding pocket is best suited for favourable host/guest interactions with N2H4. Variable temperature NMR spectra provided additional support. At 25 °C, complexes C display Cs symmetric spectra that suggest a dynamic process. Upon cooling, each undergo broadening with a coalescence temperature of Tc = 276, 260, and 240 K for C-2, C-3, and C-4, respectively. For each complex, we propose this dynamic process is the same. From the coalescence temperature of C-2 in CDCl3, we obtained an activation energy barrier for this process of 12.8 +/− 0.1 kcal/mol (see SI). This value is similar to a previously reported on/off binding event between Zn and a ligand-tethered amine (13 kcal/mol).17 Across the series of compounds C, this energy varies by ~ 1.5 kcal/mol. Complex C-4 displays both the lowest barrier for activation (11.5 kcal/mol), as well as the longest Zn-N2H4 bond distance.

Two limiting dynamic acid/base interactions are possible in complexes C: 1) Zn-N2H4 bond scission, or 2) R3B-N2H4 bond scission. The latter was probed by attempting to form a borane- free analogue of complexes C. Treating (butylNNtBu)ZnBr2, a borane-free surrogate where the alkyl-BBN portion of the ligand was replaced with n-butyl, N2H4 caused immediate demetalation of the ligand. These results suggest that boron-nitrogen bond scission in complexes C may result in decomposition, and highlights the requirement of an appended Lewis acid for stability. To ascertain the differences in Lewis acid strengths, we measured solution Gutmann-Beckett acidities.18 These data show that the Zn in C-2 is significantly more acidic than the appended borane (acceptor number = 66.3 vs. 25.0; see SI). The acceptor numbers suggest that a competitive base may promote dissociation of the weaker acid.19 Treating compounds C with 1.5 equiv. [Bu4N][Br] rapidly regenerated compounds B, highlighting the lability of the Zn-N2H4 bond (Fig. 3A).20 This result highlights the challenges of experimentally measuring and comparing Lewis acidities: due to hard/soft-acid/base mismatches and steric considerations, measured acidities are substrate specific and do not always correspond to accessible acidities.21

We propose the dynamic solution behaviour for complexes C is the result of dissociation of the Zn-N2H4 bond (Fig. 4, top). Thermodynamically, the tether length has minimal effect on the Zn-N2H4 bond dissociation energy.22 We computationally probed the electron density distribution of the Zn-N2H4 unit via density functional theory (DFT) methods. Complexes C were analysed at the B3LYP/6–311+G(2d,p) (CH2Cl2) level of theory for all atoms except Zn (6–311+G(2d) level). Natural bond orbital analyses are consistent with the extracted thermodynamic parameters from NMR spectroscopy: changing tether length results in minimal variation to the natural charges or bond indices. Kinetically, the tether length has a clear effect. Measured rates of Zn-N2H4 on/off binding at the coalescence temperature, kc, for each complex follows the trend C-2 > C-3 > C-4.23 This trend can be rationalized in terms of the distances between the two acidic residues, Zn and boron, in complexes B and C. As the tether length is increased, the rate of on/off binding is slowed because the distance between the two acids increases.

Fig. 4.

Fig. 4

A) Top: Dynamic solution behaviour of complexes C. Left: reversible deprotonation of C-3 to form D-3 and molecular structure of D-3 (50% probability ellipsoids, only H-atoms attached to heteroatoms displayed). B) Calculated pKa values.

The acidification of hydrazine by Lewis acids was probed through DFT assessment of C-3, N2H4, and its adduct with 9-methyl-9-borabicyclo[3.3.1]nonane (9-Me-9-BBN).24 Upon coordination of N2H4 to 9-Me-9-BBN, the proximal N-H protons are acidified by 20 pKa units (Fig 4B). In C-3, the N-H protons are further acidified by ca. 10 pKa units. Notably, the Lewis acidic Zn in C-3 acidifies the N-H protons to a greater extent (pKa = 24.9 vs. 26.9; Fig. 4B, right) than the trialkylborane, supporting the measured acceptor numbers.25 Overall, these data indicate that the addition of both Zn and boron Lewis acids serve additive roles to increase the acidity of coordinated N2H4.

We assessed the viability of deprotonating the cooperatively captured N2H4. Treating a THF solution of C-3 with KN(SiMe3)2 at low temperature resulted in formation of HN(SiMe3)2 and a C1 symmetric complex by 1H NMR spectroscopy—consistent with formation of a [N2H3]1− moiety and production of (3-BBNNNtBu)ZnBr(N2H3) (D-3).26 In contrast, when similar reactions were attempted for the other tether lengths to form D-2 and D-4, we observed either demetalation or an intractable mixture (see SI), illustrating a unique stabilizing effect for the 3-carbon variant.

Due to the multiple bonding modes possible with [N2H3]1−,27 the solid-state structure was determined by SC-XRD. D-3 represents the first structurally characterized example of a Zn hydrazido1− complex. Both boron and Zn are attached to the same nitrogen of the hydrazido ligand (Fig. 4, bottom). Deprotonation results in a decrease in both the Zn-N (1.973(2); Δ = 0.074 Å) and B-N (1.630(4); Δ = 0.027 Å) bond lengths, compared to C-3. In D-3, the terminal -NH2 does not interact with Lewis acidic residues and the N-N distance is identical to C-3. The tandem Lewis acid/metal stabilization of [N2H3]1− in D-3, is reminiscent of the vanadium Lewis acid/base triad that employed a weakly acidic, but rigid, tetramethyl-1,3,2-dioxaborolane Lewis acid (B-N2H3 = 1.623(4) Å).28

The structure of D-3 is unique to the series because the 3-carbon tether can accommodate both μ−1,1 and μ−1,2 ligands (N2H3 and N2H4). The ability to stabilize both types of substrates enables facile rearrangement upon deprotonation of C-3 to form D-3. Importantly, this process is reversible; treating D-3 with [Ph2NH][OTf] quantitatively regenerates C-3 (Fig. 4, left). The mobility of the Lewis acid is highlighted for complexes B-3, C-3, and D-3. While operating independently, the boron atom is located 5.41 Å away from Zn (B-3). Upon forming a cooperative interaction with Zn in C-3 to capture a diatom, N2H4, the Zn-B distances decreases to 3.95 Å (Δ = 1.46 Å). Finally, following deprotonation, cooperative stabilization of the same N-atom in N2H3 decreases the Zn-B distance to 3.06 Å. Overall, the Lewis acid exhibits mobility of 2.35 Å, mirroring the distance traversed by amino acids in metalloenzymes during turnover.1a, 2

We have described a system where it is possible to probe distant-dependent substrate-Lewis acid relationships. The trialkylborane in this system was tethered by -(CH2)n- units at defined distances, from a substrate. Of note, the three-carbon tether affords the most versatility in terms of substrate accommodation. Our results suggest that while a certain tether length may provide an ideal fit for a given substrate, the versatility of the three-carbon tether may be the most useful for stabilizing a variety of high-energy reduction products of a single substrate (e.g. NxHy from N2). Further work from our lab is investigating both of these aspects.

Supplementary Material

ESI

Acknowledgments

This work was supported by the NIH (1R01GM111486-01A1 and 1R35GM136360-01). N.K.S. is a Camille Dreyfus Teacher-Scholar. J.J.K. is supported by the NIH NIGMS (F32GM126635). M.A.L. was supported through the James E. Harris Scholarship through SURP at the Univ. Michigan. The X-ray diffractometers were funded by the NSF (CHE 1625543).

Footnotes

Conflicts of interest

There are no conflicts to declare.

Electronic Supplementary Information (ESI) available: CCDC 2012063-2012075. For ESI and crystallographic data in CIF or another electronic format see DOI: 10.1039/x0xx00000x

Notes and references

  • 1.(a) Lipscomb WN; Sträter N, Chem. Rev 1996, 96, 2375–2434; [DOI] [PubMed] [Google Scholar]; (b) Marshall NM; Garner DK; Wilson TD; Gao Y-G; Robinson H; Nilges MJ; Lu Y, Nature 2009, 462, 113–116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.(a) Sawaya MR; Cannon GC; Heinhorst S; Tanaka S; Williams EB; Yeates TO; Kerfeld CA, J. Biol. Chem 2006, 281, 7546–7555; [DOI] [PubMed] [Google Scholar]; (b) Ferraroni M; Del Prete S; Vullo D; Capasso C; Supuran CT, Acta Cryst. D 2015, D71, 2449–2456; [DOI] [PubMed] [Google Scholar]; (c) Aggarwal M; Chua TK; Pinard MA; Szebenyi DM; McKenna R, Biochemistry 2015, 54, 6631–6638; [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Cronk JD; Endrizzi JA; Cronk MR; O’Neill JW; Zhang KYJ, Protein Sci. 2001, 10, 911–922. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Kang W; Lee CC; Jasniewski AJ; Ribbe MW; Hu Y, Science 2020, 368, 1381. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.(a) Matson EM; Park YJ; Fout AR, J. Am. Chem. Soc 2014, 136, 17398; [DOI] [PubMed] [Google Scholar]; (b) Ward MB; Scheitler A; Yu M; Senft L; Zillmann AS; Gorden JD; Schwartz DD; Ivanović-Burmazović I; Goldsmith CR, Nat. Chem 2018, 10, 1207. [DOI] [PubMed] [Google Scholar]
  • 5.(a) Miller AJM; Labinger JA; Bercaw JE, Organometallics 2010, 29, 4499–4516; [Google Scholar]; (b) Meng G; Lam NYS; Lucas EL; Saint-Denis TG; Verma P; Chekshin N; Yu J-Q, J. Am. Chem. Soc 2020, 142, 10571–10591; [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Talukdar K; Sinha Roy S; Amatya E; Sleeper EA; Le Magueres P; Jurss JW, Inorg. Chem 2020, 59, 6087–6099; [DOI] [PubMed] [Google Scholar]; (d) González-Fernández R; Crochet P; Cadierno V, Dalton Trans. 2020, 49, 210–222; [DOI] [PubMed] [Google Scholar]; (e) Nichols Eva M.; Derrick JS; Nistanaki SK; Smith PT; Chang CJ, Chem. Sci 2018, 9, 2952–2960. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Hale LVA; Szymczak NK, ACS Catal. 2018, 8, 6446–6461. [Google Scholar]
  • 7.(a) Kiernicki JJ; Zeller M; Szymczak NK, J. Am. Chem. Soc 2017, 139, 18194–18197; [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Kiernicki JJ; Zeller M; Szymczak NK, Inorg. Chem 2019, 58, 1147–1154; [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Kiernicki JJ; Zeller M; Szymczak NK, Inorg. Chem 2020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Kiernicki JJ; Norwine EE; Zeller M; Szymczak NK, Chem. Commun 2019, 55, 11896–11899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Kiernicki JJ; Shanahan JP; Zeller M; Szymczak NK, Chem. Sci 2019, 10, 5539–5545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Brown HC; Scouten CG; Liotta R, J. Am. Chem. Soc 1979, 101, 96–99. [Google Scholar]
  • 11. See SI for space filling model.
  • 12. Two molecules per unit cell: Zn-B = 4.687 and 4.823 Å.
  • 13. <1 equiv. N2H4 led to a dinuclear compound, see SI and Ref 8.
  • 14.Steiner T, Angew. Chem. Int. Ed 2002, 41, 48–76. [Google Scholar]
  • 15.Chen C-H; Gabbaï FP, Chem. Sci 2018, 9, 6210–6218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.(a) Yu X-Y; Cui X-B; Yang J-J; Zhang J-P; Luo Y-H; Zhang H; Gao W-P, CrystEngComm 2012, 14, 4719–4727; [Google Scholar]; (b) Wang Y-N; Bai F-Q; Yu J-H; Xu J-Q, Dalton Trans. 2013, 42, 16547–16555. [DOI] [PubMed] [Google Scholar]
  • 17.He Z; Craig DC; Colbran SB, J. Chem. Soc., Dalton Trans 2002, 4224–4235. [Google Scholar]
  • 18.Beckett MA; Strickland GC; Holland JR; Sukumar Varma K, Polymer 1996, 37, 4629–4631. [Google Scholar]
  • 19. Hard/soft mismatches complicate acidity comparisons.
  • 20. Yield is quantitative for C-3/C-4 and 60% for C-2, see SI.
  • 21.(a) Plumley JA; Evanseck JD, J. Phys. Chem. A 2009, 113, 5985–5992; [DOI] [PubMed] [Google Scholar]; (b) Sivaev IB; Bregadze VI, Coord. Chem. Rev 2014, 270–271, 75–88. [Google Scholar]
  • 22. BDE values from single point calc. derived from X-ray data.
  • 23. There is error in measuring peak separation in the slow-exchange limit and these values should be viewed as lower limits. NMR simulations have been unsuccessful.
  • 24. Of C, only C-3 can be successfully deprotonated. See below.
  • 25. Equilibrium conditions cannot be met to experimentally.
  • 26. For bulk synthesis, deprotonation of B-3 is preferred, see SI.
  • 27.Li Y; Shi Y; Odom AL, J. Am. Chem. Soc 2004, 126, 1794–1803. [DOI] [PubMed] [Google Scholar]
  • 28.Tutusaus O; Ni C; Szymczak NK, J. Am. Chem. Soc 2013, 135, 3403–3406. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ESI

RESOURCES