Next Article in Journal
Physicochemical and Functional Properties of Okra Leaf Polysaccharides Extracted at Different pHs
Next Article in Special Issue
Influence of the Support Composition on the Activity of Cobalt Catalysts Supported on Hydrotalcite-Derived Mg-Al Mixed Oxides in Ammonia Synthesis
Previous Article in Journal
In Situ FBG Monitoring of a Henequen-Epoxy Biocomposite: From Manufacturing to Performance
Previous Article in Special Issue
Wasteless Synthesis and Properties of Highly Dispersed MgAl2O4 Based on Product of Thermal Activation of Gibbsite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Assembly of a 3D Cobalt(II) Supramolecular Framework and Its Applications in Hydrofunctionalization of Ketones and Aldehydes

1
Department of Sciences, John Jay College and PhD Program in Chemistry, The Graduate Center, The City University of New York, New York, NY 10019, USA
2
Department of Chemistry, Hunter College, The City University of New York, New York, NY 10065, USA
*
Author to whom correspondence should be addressed.
Current address: Syosset High School, 70 S Woods Rd., Syosset, NY 11791, USA.
Chemistry 2022, 4(2), 393-404; https://doi.org/10.3390/chemistry4020029
Submission received: 7 March 2022 / Revised: 31 March 2022 / Accepted: 26 April 2022 / Published: 28 April 2022
(This article belongs to the Special Issue Earth-Abundant Metal Chemistry and Catalysis)

Abstract

:
A ditopic nitrogen ligand (E)-N′-(pyridin-4-ylmethylene)isonicotinohydrazide (L) containing both divergent pyridyl coordination sites and a hydrogen-bonding hydrazide–hydrazone moiety was synthesized. The Co(NCS)2-mediated self-assembly of L has resulted in the synthesis of a novel 3-dimensional (3D) supramolecular framework (1) that features both coordination and hydrogen bonding interactions. X-ray structural analysis reveals the formation and coordination mode of 1 in the solid state. The rational utilization of coordination bonds and hydrogen bonding interactions is confirmed and responsible for constructing the 3D materials. Catalytic studies using 1 in the presence of an activator are performed for the hydroboration and hydrosilylation reactions of ketones and aldehydes, and the results are compared with previously reported cobalt-based polymeric catalysts.

Graphical Abstract

1. Introduction

The design and synthesis of metal–organic materials with well-defined structural topology and intriguing properties are attracting tremendous attention in recent years [1,2,3]. Among the driving forces responsible for the formation of novel materials between organic ligands and metal ions, the coordination and hydrogen-bonding interactions are considered to be the most important. In general, robust 1D or 2D networks can be first constructed by strong metal–organic coordination bonds and then higher dimensional structures will be assembled by hydrogen-bonding interactions [4]. Indeed, such a concept has been implemented in the construction of a number of functional metal–organic framework materials that have displayed fascinating properties in gas absorption/storage and catalysis in recent decades [5,6,7,8].
Metal-catalyzed hydroboration of unsaturated organic functionalities is one of the most attractive methods for the synthesis of alkylboronate esters that could be readily converted into the corresponding alcohols or used for C-C bond coupling reactions [9]. Although stoichiometric methods existed for a long time for the hydroboration of some functional groups such as carbonyls, nitrile, alkenes and alkynes by borohydride reagents, chemoselective hydroboration has been a challenge [10,11]. Except for the earlier contributions using noble metal catalysts, recent progress has focused on nonprecious, earth-abundant metals, such as Fe, Co, Ni, Cu, Zn, Mn, V, Al, Mg, etc., and thus far, numerous well-defined molecular complexes based on such base metals have been reported, including those showing good chemo- and regio-selectivity [9,12,13,14,15,16,17,18,19,20,21,22,23,24]. In addition, simplebase-catalyzed and non-catalytic hydroboration of carbonyl compounds have been recently disclosed [25,26]. It has been also reported that in the hydroboration of alkenes and alkynes some species of boranes and borohydrides have played a hidden role in the catalytic cycles [27].
We are particularly interested in the design and synthesis of base metal complexes using pincer-type PNP and NNN ligands and their applications in critical catalytic hydroboration reactions of alkenes, alkynes and carbonyl compounds [22,23,24]. Although superior catalytic performance using well-defined, discrete complexes of Co, Mn, V and Al as catalysts have been revealed, recent work has indicated that polymeric metal–organic materials assembled from ditopic pure-nitrogen ligands such as 4′-pyridyl-2,2′;6′,2″-terpyridine (tpy) and 4,2′;6′,4″-tpy upon coordinating with cobalt or iron salts can be used as highly efficient precatalysts for the hydroboration of various C = X (X = C, N or O) functionalities [28,29,30,31]. Significantly, it was revealed that the 1D polymeric chain assembled by 4′-pyridyl-2,2′;6′,2″-tpy and CoCl2 was an excellent precatalyst for the hydroboration of various carbonyl compounds, alkenes and alkynes with high turnover efficiency and notable chemo- and regioselectivity in the presence of a base activator, potassium tert-butoxide (KOtBu) [28,29,30]. In addition, a 2D coordination network of FeCl2 with a 4,2′;6′,4″-tpy derivative was found to catalyze the hydroboration of ketones and aldehydes in the air [31]. In general, such metal precatalysts utilize the rigid tpy ligands to form 1D polymeric chains or 2D coordination network, yet the synthesis of higher dimensional materials based on tpy-type ligands has been unsuccessful thus far in our group.
Intrigued by the fascinating catalytic properties of 3D metal–organic framework materials assembled by both coordination and hydrogen bonds, we continued on our catalyst design by moving from tpy ligands to semi-rigid Schiff-based ligands that contain both pyridine coordination sites and hydrogen bonding moiety. Thus, the structurally simple and synthetically facile ligand, L (Scheme 1), was designed in order to obtain novel metal–organic framework structures upon coordinating with octahedral metal ions such as cobalt(II). We have previously studied the fundamental coordination chemistry of similar ligands containing a hydrazide–hydrazone structural unit that can potentially form hydrogen bonding networks [32,33]. Herein, we report the synthesis and structural characterization of a 3D supramolecular framework (1) assembled by L and Co(NCS)2 and its catalytic applications in chemoselective hydroboration and hydrosilylation of ketones and aldehydes.

2. Materials and Methods

2.1. Materials

Anhydrous grade solvents and liquid reagents used were obtained from Sigma-Aldrich (Milwaukee, WI, USA)and Fisher Scientific (Pittsburgh, PA, USA) and stored over 4 Å molecular sieves. FT-IR spectra were recorded on a Shimadzu 8400S instrument (Shimadzu Scientific Instruments, Columbia, SC, USA) with solid samples under N2 using a Golden Gate ATR accessory. Elemental analysis was performed by the Midwest Microlab LLC in Indianapolis, IN, USA. Powder X-ray diffraction (PXRD) was performed with a Bruker D8 Discover microdiffractometer (Bruker Scientific LLC, Billerica, MA, USA) with the General Area Detector Diffraction System (GADDS) equipped with a VANTEC-2000 2D detector. GC-MS analysis was obtained using a Shimadzu GCMS-QP2010S gas chromatograph mass spectrometer. Ligand L was prepared according to the literature [34,35].

2.2. Synthesis of 1

A solution of (E)-N′-(pyridin-4-ylmethylene)isonicotinohydrazide (L, 113 mg, 0.5 mmol) in CH2Cl2/MeOH (2:1, v/v, 15 mL) was placed in a test tube. MeOH (5.0 mL) was layered carefully on the top of the first solution, followed by a layer of a solution of Co(SCN)2 (87.5 mg, 0.5 mmol) in MeOH (10 mL). The tube was capped and allowed to stand for one week at room temperature, during which period pale pink crystals suitable for single-crystal X-ray diffraction had formed on the walls of the tube and were isolated by decanting the solvent. The crystals were washed with MeOH and then dried in the air. Yield: 134 mg (86% based on L). FT-IR (solid, cm−1): 3207 m, 2049 s, 1657 s, 1613 s, 1548 s, 1493 w, 1293 s, 1223 m, 1147 w, 1115 w, 1066 s, 1014 s, 948 m, 816 s, 657 s. Anal. Calc. for C26H20CoN10O2S2·CH3OH, C 49.17, H 3.67, N 21.24%; Found C 49.07, H 3.99, N 20.95%.

2.3. General Procedure for 1-Catalyzed Hydroboration of Ketones and Aldehydes

In a glovebox under a nitrogen atmosphere, 1 (0.63 mg, 1.0 μmol, based on Co(L)2(NCS)2) and KOtBu (1.1 mg, 10 μmol) were placed in a 1.8 mL disposable vial equipped with a stir bar. Ketone or aldehyde (1.0 mmol) and pinacolborane (HBpin, 140.8 mg, 1.1 mmol) were then added. The reaction mixture was allowed to stir at room temperature for 2 h. At completion of the reaction, the reaction was exposed to the air and diethyl ether (10 mL) was added. The crude mixture was treated with 2N NaOH (1 mL) and 30% H2O2 (1 mL) stirred at room temperature for 1 h. The solution was extracted with ethyl acetate and washed with brine and water. The organic phase was concentrated under reduced pressure and then purified through a flash column chromatography with SiO2 using ethyl acetate/hexane as an eluent. The products were characterized by 1H and 13C NMR spectroscopies.

2.4. General Procedure for 1-Catalyzed Hydrosilylation of Ketones and Aldehydes

In a glovebox under a nitrogen atmosphere, 1 (0.63 mg, 1.0 μmol, based on Co(L)2(NCS)2) and KOtBu (1.1 mg, 10 μmol) were placed in a 1.8 mL disposable vial equipped with a stir bar. Ketone or aldehyde (1.0 mmol) and phenylsilane (PhSiH3, 119.0 mg, 1.1 mmol) were then added. The reaction mixture was allowed to stir at room temperature for 2 h. At completion of the reaction, the reaction was exposed to the air and diethyl ether (10 mL) was added. The crude mixture was treated with 2N NaOH (1 mL) at room temperature for 2 h. The solution was extracted with ethyl acetate and washed with brine and water. The organic phase was concentrated under reduced pressure and then purified through a flash column chromatography with SiO2 using ethyl acetate/hexane as an eluent. The products were characterized by 1H and 13C NMR spectroscopies.

2.5. General Procedure for 1-Catalyzed Chemoselective Hydroboration

In a glovebox under a nitrogen atmosphere, 1 (0.63 mg, 1.0 μmol, based on Co(L)2(NCS)2) and KOtBu (1.1 mg, 10 μmol) were placed in a 1.8 mL disposable vial equipped with a stir bar. Ketone (1.0 mmol), another reducible substrate (1.0 mmol) and HBpin (128 mg, 1.0 mmol) were then added. The reaction mixture was allowed to stir at room temperature for 2 h. At completion of the reaction, the reaction was exposed to the air and the solvent was evaporated. The hydroborated product was analyzed by GC-MS using hexamethylbenezene as an internal standard to determine the yield.

2.6. X-ray Crystallography

A grey-pink crystal of 1 was mounted on a Cryoloop with Paratone-N oil and data were collected at 100 K with a Bruker APEX II CCD using Mo Kα radiation generated from a rotating anode. Data were corrected for absorption with SADABS (Version 2012/1) and the structure was solved by direct methods. All non-hydrogen atoms were refined anisotropically by full-matrix least-squares on F2 with SHELXL (Version 2018/3) [36,37]. Hydrogen atom H3 on nitrogen atom N3 was found from a Fourier difference map. Uiso(H) was set to 1.20 of Ueq(N), and the N3-H3 distance was refined using a DFIX command set to 0.87(1) Å (final distance refined to 0.867(10) Å). All other hydrogen atoms were placed in calculated positions with appropriate riding parameters. The unit cell contained an unknown number of CH2Cl2 and/or MeOH molecules in two separate voids in the unit cell (each was 325 Å3 with 75 electrons). These were treated as a diffuse contribution to the overall scattering without specific atom positions by SQUEEZE/PLATON [38,39]. The crystal structure was drawn with XP, and packing figures were drawn with the program Mercury v. 3.10.3. Crystallographic data for 1: C26H20CoN10O2S2, M = 627.57, grey-pink block, monoclinic, space group P21/c, a = 8.1335(12), b = 24.670(3), c = 9.4320(13) Å, β = 91.608(5), U = 1891.8(4) Å3, Z = 2, Z’ = 0.5, Dc = 1.102 Mg m–3, μ(Mo-Kα) = 0.597 mm−1, λ = 0.71073 Å, T = 100(2) K. There were 22,705 measured, 3583 unique, and 2817 observed [I > 2σ(I)] reflections. Refinement of reflections with I > 2σ(I) (190 parameters, 1 restraint) converged at R1 = 0.0395 (R1 all data = 0.0572), wR2 = 0.858 (wR2 all data = 0.0919), GOF = 1.029. CCDC No. 1978581 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via https://www.ccdc.cam.ac.uk/structures/ (accessed on 25 March 2022), or from the Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; Fax: (+44) 1223-336-033; or e-mail: deposit@ccdc.cam.ac.uk.

3. Results

3.1. Synthesis and Structural Characterization

The synthesis of ligand L was readily carried out in high yield by using a simple Schiff-based condensation between isonicotinoyl hydrazide and pyridine-4-carbaldehyde. Previously, this ligand has been utilized for the construction of mixed-ligand Zn and Cd metal–organic frameworks where the hydrazide–hydrazone moiety played an important role in fluorescent metal ion sensing and CO2 absorption [34,35]. The cobalt-directed self-assembly of L with Co(NCS)2 was conducted by a layering method in a CH2Cl2/MeOH solution. The block-like crystalline materials of 1 were obtained with an 86% yield after one week and the product was isolated by decanting the solvent and then washed with methanol and dried in the air. It was found that the large blocks of 1 tended to lose co-crystallized solvents readily upon exposure to the air and were cracked into microcrystalline materials. Single-crystal X-ray diffraction analysis was, therefore, carried out with a crystal collected rapidly from the solution and covered with paratone oil to avoid solvent loss from the lattice. Indeed, although its main structure could be unambiguously determined, there are a number of solvent molecules (CH2Cl2 and MeOH) in the unit cell that could not be well refined and hence treated with the program SQUEENE/PLATON [38,39]. The as-dried micro-crystalline sample was characterized by FT-IR and elemental analysis, as well as PXRD analysis. A comparison of the IR spectra between L and 1 revealed slight shifts of major absorptions upon coordination, except for the appearance of the characteristic absorption of the NCS group at 2053 cm−1 (see Supplementary Material). Elemental analysis confirmed the composition of 1 with an empirical formula of Co(L)2(NCS)2 containing one molecule of MeOH in each repeating unit. The PXRD measurement of the as-dried crystals revealed the partial loss of its crystallinity and a pattern distinct from that simulated from single-crystal diffraction data, indicating the possibility of a phase transition due to the partial loss of co-crystallized solvents, which are frequently observed in porous metal–organic framework [40,41,42].
In our strategy, the octahedron-coordinating cobalt(II) ions are expected to bind with the terminal pyridyl-N atoms to form a 2D supramolecular architecture, while leaving the bridging hydrazide–hydrazone area available for the further assembly by 2D coordination sheets via hydrogen bonding interactions (Scheme 1). Accordingly, X-ray structural analysis of 1 confirms that a novel 3D framework is constructed through 2D sheet-like assembly by coordination bonds between L as linkers and octahedral CoN4(NCS)2 as the nodes followed by anticipated hydrogen bonds of the hydrazide–hydrazone regions.
1 crystallizes in the monoclinic space group P2(1)/c and its ORTEP structure showing the metal coordination environment is represented in Figure 1. The asymmetric unit contains 1 independent ligand molecule, 0.5 cobalt centers and 1 thiocyanate ion. Structural expansion around the metal center leads to the formation of a hexa-coordinate CoII center. Each CoII is surrounded by four ligand molecules to form a CoN6 sphere, and the resultant Co-Npyridyl distances are 2.1760(18) and 2.2203(17) Å, while the Co-NNCS bond is slightly shorter (2.058(2) Å). The adjacent N-Co-N angles are in the range between 87.04(7) and 92.96(7)°, with the cobalt ion residing in an almost perfect octahedral coordination environment (Figure 1). Further inspection of the structure found that each ligand bonds to two cobalt centers by the end pyridyl sites, thus linking the metal centers to form a 2D sheet-like network that contains large grids (Figure 2a). In each grid structure, the adjacent Co···Co distance is 15.440 Å, while the diagonal Co···Co distances are 24.670 and 18.573 Å, respectively, due to the slight bending conformation of L. Interestingly, the 2D network was further assembled to a 3D architecture through inter-layer hydrogen bonding interactions using the hydrazide-NH as an H-bonding donor and C = O as the H-bonding acceptor (N···O = 2.827(2) Å, < N-H···O = 167(2)°), as shown in Figure 2b. Such a hydrogen bonding packing mode using the central zone of the ligands results in a more condensed 3D molecular stacking, while forming small channels in the crystals with a dimension of ca. 7 × 7 Å, which are mainly occupied by diffused solvents that rapidly evaporate in the air and can be further removed under vacuum.

3.2. Catalytic Applications

Next, we evaluated the catalytic performance of the 3D material 1 as a precatalyst for the hydroboration of ketone and aldehyde. First, acetophenone was chosen as a model substrate to react with the hydride source, pinacolborane (HBpin), in the presence of 1 and a base additive, the conditions adopted from the 1D Co coordination polymer ([Co(4-pytpy)Cl2]n)-catalyzed hydroboration as previously reported [28]. The results for the screening of reaction conditions are summarized in Table 1. It was revealed that the reaction of acetophenone with HBpin (1.1 eq.) did not proceed in the presence of 1 (0.1 mol% based on one Co(L)2(NCS)2 unit) without the addition of an activator (entry 1). Likewise, the reaction gave only a trace amount of product in the presence of KOtBu (1 mol%) alone as detected by GC-MS analysis (entry 2). In contrast, when the Co(NCS)2 salt combined with KOtBu was used, a moderate yield was observed (entry 3). The combination of 1 and KOtBu was, therefore, crucial for the formation of active catalytic species and to this end, a quantitative yield of the desired boronate ester was obtained in THF at room temperature (entry 4). The results are identical to those obtained with [Co(4-pytpy)Cl2]n as a precatalyst (entry 5). The reactivity of the present catalyst is comparable to the reported 1D cobalt coordination polymer. We further studied the effect of different solvents on this reaction (entries 6–9). It was found that although benzene and toluene were less ideal solvents, pentane and diethyl ether are viable solvents that rendered the reaction with a slightly lower yield. It was also delightfully observed that the reaction can proceed equally well without the use of a solvent (entry 10), indicating the promising applications of this catalyst in green chemistry [13,31]. We also investigated the effect of additives as it is necessary for the present catalysis. Replacing KOtBu with KOH, K2CO3, NaBF4 or NaBH4 generally resulted in lower yields of the boronate ester (entries 11–14). However, the strong base, lithium bis(trifluoromethane)sulfonimidate (LiNTf2), was found to facilitate the reaction with comparable results (entry 15).
Encouraged by the high catalytic activity found for the combined catalyst of 1/KOtBu under neat and mild conditions, we explored its application for a range of ketones and aldehydes. The results are shown in Table 2. Acetophenone bearing electro-withdrawing and -donating groups, or the bulkier phenyl isopropyl ketone, furnished the reaction with high yields of the corresponding boronate esters under neat conditions and secondary alcohols could be readily obtained by hydrolysis (entries 2–5). In addition, α, β-unsaturated ketone and heterocyclic ketone are both suitable substrates for hydroboration, affording the corresponding alcohols with 90% and 82% yields, respectively (entries 6 and 7). Furthermore, aldehydes are also found to be hydroborated effectively in the presence of 1/KOtBu under neat conditions (entries 8–10). Primary alcohols could be isolated upon hydrolysis of the boronate esters with 86–90% yields. We are also interested to see whether other hydride sources can be used for the reduction of ketones and aldehydes. Thus, the popularly used hydride, phenylsilane (PhSiH3), was tested for the relevant hydrosilylation catalysis. It was delightful to find that both ketones and aldehydes proceed well for hydrosilylation with PhSiH3 under neat conditions, and the corresponding alcohol products were isolated in 84–92% yields after hydrolysis (entries 11–15). Finally, we tried to expand the applications of this catalyst system for the hydroboration or hydrosilylation of alkenes and alkynes. Unfortunately, the results reveal that the present catalyst is almost inactive for such substrates for both types of hydrofunctionalizations (entries 16–18). This is, however, in contrast with the previously reported tpy-based cobalt(II) coordination polymer which is a highly efficient catalyst for the regioselective hydroboration of both polar and non-polar unsaturated bonds [28,29]. These results indicate the influence of the ligand on the catalytic activity/selectivity. Despite the polymeric nature of 1, the catalyst system appeared to be a homogeneous solution under the catalytic conditions (in the presence of KOtBu and HBpin) and we were unable to recycle the catalyst as a solid by filtration or centrifugation of the reaction mixture, indicating the polymeric structure might have decomposed. At this point, the hidden role played by a borane formed in the reaction mixture and borohydride could not be excluded [27].

3.3. Chemoselectivity

Chemoselectivity is a critical issue when a catalyst is considered to be used in a practical process. To establish the chemoselectivity of the present catalyst for ketone and other reducible chemicals, we conducted several additional experiments to evaluate the present catalyst. Thus, several parallel reactions were set up while using two mixed substrates under the standard catalytic conditions as shown in Scheme 2. First, equimolar acetophenone and benzaldehyde were used to react with 1 equiv. of HBpin in the presence of 1/KOtBu under neat conditions, and it was found that hydroboration selectively occurred in the aldehyde with a 90% yield, and the ketone was fully recovered. Likewise, competing experiments using equimolar acetophenone and other reducible chemicals such as alkene, alkyne, nitrile, ester and amide were performed. The results reveal that in all cases, the hydroboration was chemoselective for ketone over other functional groups. Interestingly, while the presence of other reducible substrates affected the reaction of ketone hydroboration with lower yields, the tertiary amide was found not to interrupt the reaction at all (Scheme 2).

4. Conclusions

In conclusion, we have synthesized and characterized a novel supramolecular framework (1) by the self-assembly of a pyridine-based hydrazide–hydrazone ligand (L) and Co(NCS)2. X-ray structural analysis has determined the molecular structure to be a 3D framework resulting from the hydrogen-bonding packing between 2D sheets of the extended coordination between L and metal ions. The catalytic applications of the as-synthesized polymeric materials in combination with an activator for hydroboration (and hydrosilylation) reactions of ketones and aldehydes were explored. The results suggest that 1 can be employed for both hydroboration and hydrosilylation catalysis for ketones and aldehydes under mild conditions, although it is not a viable catalyst for alkene and alkyne hydrofunctionalization. Good chemoselectivity of this catalyst system was also observed for aldehyde over ketone as well as ketone over other reducible groups such as alkene, alkyne, nitrile, ester and amide.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemistry4020029/s1, Supplementary data (catalytic details and NMR spectroscopy of isolated products). Crystallographic data for this paper have been deposited with the CCDC, accession numbers 1978581. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif (accessed on 25 March 2022) or by emailing data_request@ccdc.cam.ac.uk. Reference [43] is cited in the supplementary materials.

Author Contributions

Conceptualization, G.Z. and S.Z.; synthesis, characterization and catalysis, G.Z., A.W. and H.Z.; crystallographic analysis, M.C.N.; writing—original draft preparation, G.Z.; writing—review and editing, G.Z., A.W., S.Z. and M.C.N.; supervision, G.Z.; project administration, G.Z.; funding acquisition, G.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Science Foundation, grant number CHE-1900500 and the City University of New York PSC-CUNY awards, grant numbers 63809-0051 and 64254-0052.

Data Availability Statement

Data is contained within the article or Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Gangu, K.K.; Maddila, S.; Mukkamala, S.B.; Jonnalagadda, S.B. A Review on Contemporary Metal-Organic Framework Materials. Inorg. Chim. Acta 2016, 446, 61–74. [Google Scholar] [CrossRef]
  2. Cai, G.; Yan, P.; Zhang, L.; Zhou, H.C.; Jiang, H.L. Metal–organic framework-based hierarchically porous materials: Synthesis and applications. Chem. Rev. 2021, 121, 12278–12326. [Google Scholar] [CrossRef] [PubMed]
  3. Chakraborty, G.; Park, I.H.; Medishetty, R.; Vittal, J.J. Two-dimensional metal-organic framework materials: Synthesis, structures, properties and applications. Chem. Rev. 2021, 121, 3751–3891. [Google Scholar] [CrossRef] [PubMed]
  4. Kitagawa, S.; Uemura, K. Dynamic porous properties of coordination polymers inspired by hydrogen bonds. Chem. Soc. Rev. 2005, 34, 109–119. [Google Scholar] [CrossRef]
  5. MacDonald, J.C.; Dorrestein, P.C.; Pilley, M.M.; Foote, M.M.; Lundburg, J.L.; Henning, R.W.; Schultz, A.J.; Manson, J.L. Design of Layered Crystalline Materials Using Coordination Chemistry and Hydrogen Bonds. J. Am. Chem. Soc. 2000, 122, 11692–11702. [Google Scholar] [CrossRef]
  6. Cui, Y.; Ngo, H.L.; White, P.S.; Lin, W. Hierarchical Assembly of Homochiral Porous Solids Using Coordination and Hydrogen Bonds. Inorg. Chem. 2003, 42, 652–654. [Google Scholar] [CrossRef]
  7. Wu, X.; Wang, J.; Huang, J.; Yang, S. Robust, Stretchable, and Self-Healable Supramolecular Elastomers Synergistically Cross-Linked by Hydrogen Bonds and Coordination Bonds. ACS Appl. Mater. Interf. 2019, 11, 7387–7396. [Google Scholar] [CrossRef]
  8. Kwon, H.; Lee, E. Coordination preference of hexa(2-pyridyl)benzene with copper(II) directed by hydrogen bonding. CrystEngComm 2018, 20, 5233–5240. [Google Scholar] [CrossRef]
  9. Chang, C.C.; Kinjo, R. Catalytic Hydroboration of Carbonyl Derivatives, Imines, and Carbon Dioxide. ACS Catal. 2015, 5, 3238–3259. [Google Scholar] [CrossRef]
  10. Cho, B.T. Recent development and improvement for boron hydride-based catalytic asymmetric reduction of unsymmetrical ketones. Chem. Soc. Rev. 2009, 38, 443–452. [Google Scholar] [CrossRef]
  11. Wang, W.; Lu, K.; Qin, Y.; Yao, W.; Yuan, D.; Pullarkat, S.A.; Xu, L.; Ma, M. Grignard reagents-catalyzed hydroboration of aldehydes and ketones. Tetrahedron 2020, 76, 131145. [Google Scholar] [CrossRef]
  12. Tamang, S.R.; Findlater, M. Emergence and Applications of Base Metals (Fe, Co, and Ni) in Hydroboration and Hydrosilylation. Molecules 2019, 24, 3194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Wang, R.; Park, S. Recent Advances in Metal-Catalyzed Asymmetric Hydroboration of Ketones. ChemCatChem 2021, 13, 1898–1919. [Google Scholar] [CrossRef]
  14. Shegavi, M.L.; Bose, S.K. Recent advances in the catalytic hydroboration of carbonyl compounds. Catal. Sci. Tech. 2019, 9, 3307–3336. [Google Scholar] [CrossRef]
  15. Jakhar, V.K.; Barman, M.K.; Nembenna, S. Aluminum monohydride catalyzed selective hydroboration of carbonyl compounds. Org. Lett. 2016, 18, 4710–4713. [Google Scholar] [CrossRef] [PubMed]
  16. Oluyadi, A.A.; Ma, S.; Muhoro, C.N. Titanocene(II)-Catalyzed Hydroboration of Carbonyl Compounds. Organometallics 2013, 32, 70–78. [Google Scholar] [CrossRef]
  17. Baishya, A.; Baruah, S.; Geetharani, K. Efficient hydroboration of carbonyls by an iron(II) amide catalyst. Dalton Trans. 2018, 47, 9231–9236. [Google Scholar] [CrossRef]
  18. Khoo, S.; Cao, J.; Ng, F.; So, C.W. Synthesis of a Base-Stabilized Silicon(I)–Iron(II) Complex for Hydroboration of Carbonyl Compounds. Inorg. Chem. 2018, 57, 12452–12455. [Google Scholar] [CrossRef]
  19. Vijjamarri, S.; O’Denius, T.M.; Yao, B.; Kubátová, A.; Du, G. Highly Selective Hydroboration of Carbonyls by a Manganese Catalyst: Insight into the Reaction Mechanism. Organometalics 2020, 39, 3375–3383. [Google Scholar] [CrossRef]
  20. Li, M.; Liu, X.; Cui, D. Catalytic hydroboration of carbonyl derivatives by using phosphinimino amide ligated magnesium complexes. Dalton Trans. 2021, 50, 13037–13041. [Google Scholar] [CrossRef]
  21. Kim, J.H.; Jaladi, A.K.; Kim, H.T.; An, D.K. Lithium tert-Butoxide-catalyzed Hydroboration of Carbonyl Compounds. Bull. Korean Chem. Soc. 2019, 40, 971–975. [Google Scholar] [CrossRef]
  22. Zhang, G.; Zeng, H.; Wu, J.; Yin, Z.; Zheng, S.; Fettinger, J.C. Highly Selective Hydroboration of Alkenes, Ketones and Aldehydes Catalyzed by a Well-Defined Manganese Complex. Angew. Chem. Int. Ed. 2016, 55, 14369–14372. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, G.; Wu, J.; Zeng, H.; Neary, M.C.; Devany, M.; Zheng, S.; Dub, P.A. Dearomatization and Functionalization of Terpyridine Ligands Leading to Unprecedented Zwitterionic Meisenheimer Aluminum Complexes and Their Use in Catalytic Hydroboration. ACS Catal. 2019, 9, 874–884. [Google Scholar] [CrossRef]
  24. Zhang, G.; Wu, J.; Zheng, S.; Neary, M.C.; Mao, J.; Flores, M.; Trovitch, R.J.; Dub, P.A. Redox-Noninnocent Ligand-Supported Vanadium Catalysts for the Chemoselective Reduction of C=X (X=O, N) Functionalities. J. Am. Chem. Soc. 2019, 141, 15230–15239. [Google Scholar] [CrossRef] [PubMed]
  25. Ma, D.H.; Jaladi, A.K.; Lee, J.H.; Kim, T.S.; Shin, W.K.; Hwang, H.; An, D.K. Catalytic Hydroboration of Aldehydes, Ketones, and Alkenes Using Potassium Carbonate: A Small Key to Big Transformation. ACS Omega 2019, 4, 15893–15903. [Google Scholar] [CrossRef] [Green Version]
  26. Nowicki, M.; Kuciński, K.; Hreczycho, G.; Hoffmann, M. Catalytic and non-catalytic hydroboration of carbonyls: Quantum-chemical studies. Org. Biomol. Chem. 2021, 19, 3004–3015. [Google Scholar] [CrossRef]
  27. Bage, A.D.; Nicholson, K.; Hunt, T.A.; Langer, T.; Thomas, S.P. The Hidden Role of Boranes and Borohydrides in Hydroboration Catalysis. ACS Catal. 2020, 10, 13479–13486. [Google Scholar] [CrossRef]
  28. Wu, J.; Zeng, H.S.; Cheng, J.; Zheng, S.P.; Golen, J.A.; Manke, D.R.; Zhang, G. Cobalt(II) coordination polymer as a precatalyst for selective hydroboration of aldehydes, ketones, and imines. J. Org. Chem. 2018, 83, 9442–9448. [Google Scholar] [CrossRef]
  29. Zhang, G.; Wu, J.; Li, S.; Cass, S.; Zheng, S. Markovnikov-Selective Hydroboration of Vinylarenes Catalyzed by a Cobalt (II) Coordination Polymer. Org. Lett. 2018, 20, 7893–7897. [Google Scholar] [CrossRef]
  30. Zhang, G.; Li, S.; Wu, J.; Zeng, H.; Mo, Z.; Davis, K.; Zheng, S. Highly efficient and selective hydroboration of terminal and internal alkynes catalysed by a cobalt(II) coordination polymer. Org. Chem. Front. 2019, 6, 3228–3233. [Google Scholar] [CrossRef]
  31. Zhang, G.; Cheng, J.; Davis, K.; Bonifacio, M.G.; Zajaczkowski, C. Practical and selective hydroboration of aldehydes and ketones in air catalysed by an iron(II) coordination polymer. Green Chem. 2019, 21, 1114–1121. [Google Scholar] [CrossRef]
  32. Liu, E.; Li, L.; Cheng, J.; Zhang, G. Synthesis and structural characterization of dinuclear Zinc(II) and Europium(III) complexes based on a bis-hydrazone ligand. J. Mol. Struct. 2019, 1188, 1–6. [Google Scholar] [CrossRef]
  33. Li, L.; Zhang, Y.Z.; Yang, C.; Liu, E.; Golen, J.A.; Rheingold, A.L.; Zhang, G. Synthesis and structural characterization of zinc(II) and cobalt(II) complexes based on multidentate hydrazone ligands. J. Mol. Struct. 2016, 1110, 180–184. [Google Scholar] [CrossRef] [Green Version]
  34. Parmar, B.; Rachuri, Y.; Bisht, K.K.; Suresh, E. Mixed-Ligand LMOF Fluorosensors for Detection of Cr(VI) Oxyanions and Fe3+/Pd2+ Cations in Aqueous Media. Inorg. Chem. 2017, 56, 10939–10949. [Google Scholar] [CrossRef]
  35. Roztocki, K.; Jędrzejowski, D.; Hodorowicz, M.; Senkovska, I.; Kaskel, S.; Matoga, D. Effect of Linker Substituent on Layers Arrangement, Stability, and Sorption of Zn-Isophthalate/Acylhydrazone Frameworks. Cryst. Growth Des. 2018, 18, 488–497. [Google Scholar] [CrossRef]
  36. Sheldrick, G.M. SHELXTL, An Integrated System for Solving, Refining, and Displaying Crystal Structures from Diffraction Data; University of Göttingen: Göttingen, Germany, 1981. [Google Scholar]
  37. Sheldrick, G.M. Crystal Structure Refinement by SHELXL. Acta Cryst. 2015, 71, 3–8. [Google Scholar]
  38. Spek, A.L. PLATON, A Multipurpose Crystallographic Tool; Utrecht University: Utrecht, The Netherlands, 2005. [Google Scholar]
  39. Spek, A.L. Single-Crystal Structure Validation with the Program PLATON. J. Appl. Cryst. 2003, 36, 7–13. [Google Scholar] [CrossRef] [Green Version]
  40. Liu, X.L.; Fan, W.W.; Lu, Z.X.; Qin, Y.; Yang, S.X.; Li, Y.; Liu, Y.X.; Zheng, L.Y.; Cao, Q.E. Solvent-Driven Reversible Phase Transition of a Pillared Metal–Organic Framework. Chem. Eur. J. 2019, 25, 5787–5792. [Google Scholar] [CrossRef]
  41. Krūkle-Bērziņa, K.; Belyakov, S.; Mishnev, A.; Shubin, K. Stability and Phase Transitions of Nontoxic γ-Cyclodextrin-K+ Metal-Organic Framework in Various Solvents. Crystals 2020, 10, 37. [Google Scholar] [CrossRef] [Green Version]
  42. Kundu, T.; Wahiduzzaman, M.; Shah, B.B.; Maurin, G.; Zhao, D. Solvent-induced control over breathing behavior in flexible metal–organic frameworks for natural-gas delivery. Angew. Chem. Int. Ed. 2019, 58, 8073–8077. [Google Scholar] [CrossRef]
  43. Zeng, H.; Wu, J.; Li, S.; Hui, C.; Ta, A.; Cheng, S.-Y.; Zheng, S.; Zhang, G. Copper(II)-Catalyzed Selective Hydroboration of Ketones and Aldehydes. Org. Lett. 2019, 21, 401. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Schematic representation of the assembly of 2D grid-like network found in 1 from ligand L and Co(NCS)2.
Scheme 1. Schematic representation of the assembly of 2D grid-like network found in 1 from ligand L and Co(NCS)2.
Chemistry 04 00029 sch001
Figure 1. The ORTEP representation of 1 with thermal ellipsoids plotted at the 30% probability level. Hydrogen atoms bound to carbon are omitted for clarity.
Figure 1. The ORTEP representation of 1 with thermal ellipsoids plotted at the 30% probability level. Hydrogen atoms bound to carbon are omitted for clarity.
Chemistry 04 00029 g001
Figure 2. (a) The capped-stick representation of the partial 2D sheet-like coordination network. (b) The local ball-stick representation showing the interlayer N-H···O hydrogen bonding interactions drawn in green dashed lines. The grid in mid-layer is shown in blue. (c) The The 3D packing mode of 1 along the crystallographic c axis showing the void molecular channels (diffused solvents are removed for clarification).
Figure 2. (a) The capped-stick representation of the partial 2D sheet-like coordination network. (b) The local ball-stick representation showing the interlayer N-H···O hydrogen bonding interactions drawn in green dashed lines. The grid in mid-layer is shown in blue. (c) The The 3D packing mode of 1 along the crystallographic c axis showing the void molecular channels (diffused solvents are removed for clarification).
Chemistry 04 00029 g002
Scheme 2. Chemoselective hydroboration between ketone and other reducible substrates. Conditions: acetophenone (1.0 mmol), reducible substrate (1.0 mmol), 1 (0.1 mol%), KOtBu (1 mol%) and HBpin (1.0 mmol), neat, rt, 2 h.
Scheme 2. Chemoselective hydroboration between ketone and other reducible substrates. Conditions: acetophenone (1.0 mmol), reducible substrate (1.0 mmol), 1 (0.1 mol%), KOtBu (1 mol%) and HBpin (1.0 mmol), neat, rt, 2 h.
Chemistry 04 00029 sch002
Table 1. Catalytic test for 1-catalyzed hydroboration of acetophenone with HBpin a.
Table 1. Catalytic test for 1-catalyzed hydroboration of acetophenone with HBpin a.
Chemistry 04 00029 i001
EntryCatalystActivatorSolventYield/% b
11-THF0
2noneKOtBuTHF<5
3Co(NCS)2KOtBuTHF68
41KOtBuTHF99
5[Co(4-pytpy)Cl2]nKOtBuTHF99
61KOtBuEt2O95
71KOtButoluene80
81KOtBubenzene86
91KOtBupentane95
101KOtBunone99
111KOHnone58
121K2CO3none67
131NaBH4none85
141NaBF4none18
151LiNTf2none96
a Conditions: acetophenone (1.0 mmol), HBpin (1.1 mmol, 1.1 eq.), 1 (0.1 mol%), activator (1 mol%) and solvent (0.5 mL), rt, 2 h. b Determined by GC-MS analysis with hexamethylbenzene as an internal standard.
Table 2. Catalytic hydroboration or hydrosilylation of ketones and aldehydes using 1 as a precatalyst a.
Table 2. Catalytic hydroboration or hydrosilylation of ketones and aldehydes using 1 as a precatalyst a.
Chemistry 04 00029 i002
EntrySubstrateHydride Product Yield (%) b
1 Chemistry 04 00029 i003HBpin Chemistry 04 00029 i00492
2 Chemistry 04 00029 i005HBpin Chemistry 04 00029 i00694
3 Chemistry 04 00029 i007HBpin Chemistry 04 00029 i00892
4 Chemistry 04 00029 i009HBpin Chemistry 04 00029 i01095
5 Chemistry 04 00029 i011HBpin Chemistry 04 00029 i01289
6 Chemistry 04 00029 i013HBpin Chemistry 04 00029 i01490
7 Chemistry 04 00029 i015HBpin Chemistry 04 00029 i01682
8 Chemistry 04 00029 i017HBpin Chemistry 04 00029 i01888
9 Chemistry 04 00029 i019HBpin Chemistry 04 00029 i02090
10 Chemistry 04 00029 i021HBpin Chemistry 04 00029 i02286
11 Chemistry 04 00029 i023PhSiH3 Chemistry 04 00029 i02492
12 Chemistry 04 00029 i025PhSiH3 Chemistry 04 00029 i02690
13 Chemistry 04 00029 i027PhSiH3 Chemistry 04 00029 i02884
14 Chemistry 04 00029 i029PhSiH3 Chemistry 04 00029 i03085
15 Chemistry 04 00029 i031PhSiH3 Chemistry 04 00029 i03290
16 Chemistry 04 00029 i033HBpin Chemistry 04 00029 i0345 c
17 Chemistry 04 00029 i035HBpin Chemistry 04 00029 i0363 c
18 Chemistry 04 00029 i037PhSiH3 Chemistry 04 00029 i038trace c
a Conditions: substrate (1.0 mmol), HBpin or PhSiH3 (1.1 mmol, 1.1 eq.), 1 (0.1 mol%), KOtBu (1 mol%), neat, rt, 2 h. b Isolated yields. c Yields determined by GC analysis.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, G.; Wang, A.; Zeng, H.; Zheng, S.; Neary, M.C. Assembly of a 3D Cobalt(II) Supramolecular Framework and Its Applications in Hydrofunctionalization of Ketones and Aldehydes. Chemistry 2022, 4, 393-404. https://doi.org/10.3390/chemistry4020029

AMA Style

Zhang G, Wang A, Zeng H, Zheng S, Neary MC. Assembly of a 3D Cobalt(II) Supramolecular Framework and Its Applications in Hydrofunctionalization of Ketones and Aldehydes. Chemistry. 2022; 4(2):393-404. https://doi.org/10.3390/chemistry4020029

Chicago/Turabian Style

Zhang, Guoqi, Alex Wang, Haisu Zeng, Shengping Zheng, and Michelle C. Neary. 2022. "Assembly of a 3D Cobalt(II) Supramolecular Framework and Its Applications in Hydrofunctionalization of Ketones and Aldehydes" Chemistry 4, no. 2: 393-404. https://doi.org/10.3390/chemistry4020029

Article Metrics

Back to TopTop