Next Article in Journal
Synergistic Effect in Zinc Phthalocyanine—Nanoporous Gold Hybrid Materials for Enhanced Photocatalytic Oxidations
Previous Article in Journal
Wet Peroxide Oxidation of Chlorobenzenes Catalyzed by Goethite and Promoted by Hydroxylamine
Previous Article in Special Issue
Synthesis and Structure of Copper Complexes of a N6O4 Macrocyclic Ligand and Catalytic Application in Alcohol Oxidation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ni(II)-Aroylhydrazone Complexes as Catalyst Precursors Towards Efficient Solvent-Free Nitroaldol Condensation Reaction

by
Manas Sutradhar
*,
Tannistha Roy Barman
,
Armando J. L. Pombeiro
and
Luísa M.D.R.S. Martins
*
Centro de Química Estrutural, Instituto Superior Técnico, Universidade de Lisboa, Av. Rovisco Pais, 1049-001 Lisboa, Portugal
*
Authors to whom correspondence should be addressed.
Catalysts 2019, 9(6), 554; https://doi.org/10.3390/catal9060554
Submission received: 9 June 2019 / Revised: 17 June 2019 / Accepted: 17 June 2019 / Published: 20 June 2019
(This article belongs to the Special Issue Recent Advances in Homogeneous Catalysis)

Abstract

:
The aroylhydrazone Schiff bases 2-hydroxy-(2-hydroxybenzylidene)benzohydrazide and (2,3-dihydroxybenzylidene)-2-hydroxybenzohydrazide have been used to synthesize the bi- and tri-nuclear Ni(II) complexes [Ni2(L1)2(MeOH)4] (1) and [Ni3(HL2)2(CH3OH)8]· (NO3)2 (2). Both complexes have been characterized by elemental analysis, spectroscopic techniques [IR spectroscopy and electrospray ionization-mass spectrometry (ESI-MS)], and single-crystal X-ray crystallography. The coordination behavior of the two ligands is different in the complexes: The ligand exhibits the keto form in 2, while coordination through enol form was found in 1. Herein, the catalytic activity of 1 and 2 has been compared with the nitroaldol condensation reaction under various conditions. Complex 2 exhibits the highest activity towards solvent-free conditions.

Graphical Abstract

1. Introduction

The nitroaldol (Henry) reaction, a very useful and significant reaction, plays a vital role in the synthetic organic chemistry. It is widely used to build C–C bonds by the formation of β-nitroalcohols derived from coupling of a carbonyl compound with an alkyl nitro one with α- hydrogen atoms. The reaction takes place via construction of asymmetric centers (one or two) through the new C–C junction leading to optically active products. Inorganic metal complexes or inorganic/organic bases can catalyze this reaction [1,2,3,4,5,6,7,8,9,10]. The prepared functionalized β-nitroalcohols are known as important synthetic precursors, can be converted by nucleophilic displacement into other functionalities like α-hydroxy ketones, aldehydes, carboxylic acids, azides, sulphides, etc., of pharmaceutical significance [5,6,7,8], polyfunctionalized materials [11], and are used to generate compounds of biological importance such as 1,2-diaminoalcohol, [12] aminosugars, [13] nitroketones, [14] α,β-unsaturated nitro compounds [15], and many other important bifunctional compounds. The stereoselectivity nature of the Henry reaction was identified for the first time by Shibasaki et al. [16] using several chiral metal complexes [17,18,19,20,21,22,23,24,25,26,27,28,29] and chiral organocatalysts [20], exhibiting the growing interest in this field. The reaction has been studied using different conditions like homogeneous [5,6,7,8], heterogeneous [21,22,23], ionic liquids or supercritical media [24], materials of mesoporous nanocomposite [25], etc.
The design of the catalyst plays a crucial role in controlling the diastereo- and enantioselectivity of the products, leading to a serious task for researchers to search for efficient and selective catalysts from synthetic, economic, and environmental perspectives. Di and polynuclear metal complexes including coordination polymers can efficiently catalyze the nitroaldol reaction between aldehydes and nitroalkanes [26,27,28]. Solvent-free organic synthesis has received significant interest from the chemist worldwide due to its advantage over chemical wastes in terms of sustainability and the requirements of green chemistry [29]. The use of solvent-free condition is also applied in the nitroaldol reaction [30,31]. Some nickel complexes showed good catalytic activity towards nitroaldol reactions under various conditions such as in homogeneous reaction states [32,33], under solvent-free microwave irradiation [25], in ionic liquids [34,35], or under heterogeneous [36,37,38] conditions. Aroylhydrazone Schiff bases form highly stable complexes with transition metals in various oxidation states, coordination numbers, and nuclearities, and can be tuned easily by changing different carbonyl derivatives [39,40,41,42,43,44,45,46].
In this study, we describe the synthesis and characterization of one dinuclear and one trinuclear Ni(II) complex derived from two different aroylhydrazones, and their activity as catalyst precursors towards nitroaldol (Henry) reactions to achieve desired functionalized products under different homogeneous catalytic reactions conditions. Solvent-free conditions are found most efficient in our catalytic system. This is probably due to more accessibility of substrate molecules to the metal centre under solvent-free condition than the presence of solvent molecule; 94%–97% yields are observed in our case, which is relatively higher than the yield found in other di or polynuclear catalytic system [26].

2. Results and Discussion

2.1. Syntheses and Characterizations

In this study, we have used two different aroylhydrazone Schiff bases, namely, 2-hydroxy-(2-hydroxybenzylidene)benzohydrazide (H2L1) and (2,3-dihydroxybenzylidene)-2-hydroxybenzohydrazide (H3L2) [47,48], to synthesize two different (one dinuclear and one trinuclear) Ni(II) complexes. In the dinuclear complex [Ni2(L1)2(MeOH)4] (1), the ligand exhibits the enol form with the loss of all (two) acidic hydrogens whereas H3L2 undergoes coordination with the loss of two acidic hydrogens (out of three) and one remains protonated to form the trinuclear [Ni3(HL2)2(CH3OH)8]·(NO3)2 (2) (Scheme 1). In both complexes, the phenolate oxygen at the ortho- position (from the aldehyde moiety of the ligand) forms a phenoxido bridge between two Ni(II) ions. Characterizations of 1 and 2 have been carried out by elemental analysis, spectroscopic methods (IR spectroscopy, ESI-MS), and X-ray diffraction (single crystal) techniques. Beside similar characteristic stretching signals of the ligand, a band at 1609 cm−1 appears in the IR spectrum of 2 that corresponds to the C=O stretching frequency [47,48]. The m/z value of 2 suggests the loss of two noncoordinate nitrate ions and one acidic proton from one of the two ligands present in 2 (see Experimental). The catalytic properties of 1 and 2 were investigated towards solvent-free nitroaldol condensation reaction and their activities were compared.

2.2. General Description of the Crystal Structures

Crystals of [Ni2(L1)2(DMF)4] [Ni2(L1)2(DMF)2(H2O)2]·2DMF (1A) suitable for X-ray diffractions were obtained upon re-crystallization of 1 from DMF-water and slow evaporation of 2 from the methanolic solution, at ambient temperature. The crystallographic data are summarized in Table 1, representative molecular structures are displayed in Figure 1 and Figure 2, and selected dimensions are presented in Table 2.
The asymmetric unit of the compound 1A comprises a half unit of two different molecules and one solvent DMF. One half unit of one molecule contains the nickel(II) cation with one coordinated ligand, one DMF, and one water molecule. Another half consists of nickel(II) cation with one coordinated ligand and two DMF molecules. All the Ni(II) centers exhibit a distorted octahedral coordination environment. The structure of 1A contains crystallographically generated inversion centres in the middle of the Ni1–Ni1ior Ni2–Ni2ii bonds, therefore in the heart of the respective Ni2O2 planes. The aroylhydrazones act as dianionic and tetradentate ONOO chelating equatorial ligands, binding to one of the Ni(II) centers via the enolate oxygen, the imino nitrogen, and the deprotonated phenolate oxygen, which is connected to other metal cation. Thus, the structure displays two µ-O bridges in each molecular unit, which connect the two nickel(II) centers. In the axial positions, there are two DMF oxygens atoms in one unit and DMF and water in the other unit. The N−N bond distances of 1.385 (3) and 1.397 (3) Å for the coordinated ligand indicate their single bond hydrazino character. The ODMF–Ni–ODMF or ODMF–Ni–Owater groups are nearly linear (170.47 (9)° and 172.52 (7)°) and in the Ni2(μ-O)2 cores, the Ni–O–Ni angles are ca. 100°. The Ni–Ni contact distances are about 3.10 Å.
Compound 2 crystallizes in the triclinic P¯1 space group. Its unit cell contains one molecule of the compound and two nitrate anions. Compound 2 is trinuclear with H3L2 coordinating to the Ni(II) cations in the dianionic (HL2)2− form using both phenolate O atoms, the keto O atom, and the imine N atom. The phenolate oxygen at the ortho position of the aldehyde moiety exhibits µ-O bridges with the other metal cation, which is located at the inversion centre. The asymmetric units of 2 contains half of the molecules, i.e., one and a half nickel cations, one (HL2)2− ligand, four methanol molecules, and one nitrate anion. All the axial positions are occupied by methanol molecules. The terminal nickel cations exhibit distorted octahedral N1O5 coordination environment but the Ni(II) at center of inversion displays more regular octahedral geometry. The Ni–Ni contact distances are 3.764 Å, longer than the distances found in 1A.

2.3. Catalytic Studies

The catalytic activity of complexes 1 and 2 for the nitroaldol (or Henry) reaction was evaluated using benzaldehydes and nitroethane as models (Scheme 2, Table 3 and Table 4). An excess nitroethane was used to obtain maximum conversion of benzaldehydes into products.
The selectivity for the β-nitroethanol products was 100% for all the experiments (the only compounds found besides syn- and anti- β-nitroethanols were non-reacted substrates).
Trinuclear complex 2 showed a higher activity than the dinuclear one, at the same temperature and under similar conditions (Table 3). Consequently, the reaction conditions (temperature, reaction time, amount of catalyst, and type of solvent) were optimized using a model nitroethane-benzaldehyde system with catalyst 2 (Table 3). Thus, the optimal experimental conditions found are exhibited in entry 10 of Table 3, leading to a β-nitroethanol yield of 94% (TOF = 3.9 h−1) and a good selectivity syn:anti (77:23).
A blank test was performed with neat benzaldehyde and nitroethane in the absence of any metal catalyst, at 60 °C. No β -nitroalkanol was detected after 24 h of reaction time (entry 12, Table 3). The nitroaldol reaction also did not occur appreciably (9% yield of nitroethanol) by using Ni(OAc)2 instead of the catalyst precursor 1 or 2 (entry 13, Table 3).
To determine the catalytic performance of 2 in the nitroaldol condensation reaction, a series of aromatic and aliphatic aldehydes were selected to screen as starting materials (Table 4). In general, the reactivity of the substituted benzaldehydes is less than benzaldehyde itself, possibly the presence of steric hindrance. In the case of substituted aromatic aldehydes having electron-withdrawing substituents in para-position, higher yields (entries 5 and 7, Table 4) are obtained, which may be attributed to the higher electrophilicity of the substrate compared to the aldehydes bearing electron-donating moieties. A maximum yield of 97% with a syn: anti diastereoselectivity ratio of 80:20 was observed for the aldehyde, which is para-nitro-substituted (entry 7, Table 4). For the aliphatic aldehydes, acetaldehyde and propionaldehyde (entries 8 and 9, Table 4), a maximum of 78.2% yield was obtained for propionaldehyde (entry 9, Table 4) with a syn: anti diastereoselectivity ratio of 74:26.
The catalytic activity of compound 2 was also compared in the reactions using various substituted aromatic aldehydes with different nitroalkanes, producing the corresponding β-nitroalkanols (Scheme 3), leading to yields ranging from 38% to 94% (Table 5).
In general, the yield of β-nitroalkanol decreased in the order of nitromethane > nitroethane > 1-nitropropane for both aldehydes, and the same molecular size dependent behavior was found for the aldehydes (Table 5). A proposed catalytic cycle promoted by Ni(II) centre is presented in Scheme 4 showing the reaction pathway towards the formation of the β-nitroalkanols.

3. Materials and Methods

The syntheses for this study were performed in air and reagents and solvents (commercially available) that were used as received, without further purification. The metal source for the synthesis of complexes was Ni(NO3)2·6H2O. Elemental analyses (C, H, and N) were carried out by the Microanalytical Service of the Instituto Superior Técnico. Infrared spectra (4000–400 cm−1) were recorded on a Bruker Vertex 70 instrument in KBr pellets (Bruker Corporation, Ettlingen, Germany); wavenumbers are in cm−1. The 1H NMR spectra were recorded on a Bruker Avance II + 400.13 MHz (UltraShieldTM Magnet) spectrometer at room temperature. The internal reference was tetramethylsilane. The chemical shifts are reported in ppm in the 1H NMR spectra. Mass spectra were recorded in a Varian 500-MS LC Ion Trap Mass Spectrometer (Agilent Technologies, Amstelveen, The Netherlands) equipped with an electrospray (ESI) ion source. For electrospray ionization, the drying gas and flow rate were optimized according to the particular sample with 35 p.s.i. nebulizer pressure. Scanning was carried out from m/z 100 to 1200 in methanol solution. The compounds were observed in the positive and negative mode (capillary voltage = 80–105 V).

3.1. Syntheses of the Pro-Ligand H2L

The aroylhydrazone pro-ligand 2-hydroxy-(2-hydroxybenzylidene)benzohydrazide (H2L1) and (2,3-dihydroxybenzylidene)-2-hydroxybenzohydrazide (H3L2) (Scheme 1) were prepared by a reported method [47,48] upon condensation of the 2-hydroxybenzohydrazine with 2-hydroxybenzaldehyde or 2,3-dihydroxybenzaldehyde.

3.2. Synthesis of the Nickel(II) Complexes

  • Synthesis of [Ni2(L1)2(MeOH)4] (1)
First, 0.26 g (1.01 mmol) of H2L1 was dissolved in 25 mL of methanol, and Ni(NO3)2·6H2O (0.32 g, 1.10 mmol) were added to it. The resultant mixture was stirred for 15 min at 50 °C; a dark green solution was obtained. The dark green solution was then filtered, and the solvent was allowed to evaporate slowly. After 1 d, a green microcrystalline powder was obtained, washed 3 times with cold methanol, and dried in open air.
Yield: 0.293 g (78%, with respect to Ni(II)). Anal. Calcd for (1) C32H36N4Ni2O10: C, 50.97; H, 4.81; N, 7.43. Found: C, 50.93; H, 4.78; N, 7.39. IR (KBr; cm1): 3416 ν(OH), 1609 ν(C=N), 1254 ν(C–O) enolic and 1154 ν(N–N). ESI-MS(+): m/z 753 [1+H]+ (100%).
  • Synthesis of [Ni2(L1)2(DMF)4] [Ni2(L1)2(DMF)2(H2O)2]·2DMF (1A)
The green microcrystalline powder of 1 was dissolved in DMF and few drops of water was added. After ca. 3 d, nice green crystals were isolated from the solution. Isolated compound was washed 3 times with cold methanol and kept in open air for drying. The crystals are insoluble in common organic solvents.
Anal. Calcd for (1A) C80H96N16Ni4O22: C, 51.42; H, 5.18; N, 11.99. Found: C, 51.39; H, 5.16; N, 11.93. IR (KBr; cm1): 3403 ν(OH), 1608 ν(C = N), 1258 ν(C–O) enolic and 1157 ν(N–N).
  • Synthesis of [Ni3(HL2)2(CH3OH)8]·(NO3)2 (2)
First, 0.28 g, 1.02 mmol of H3L2 was dissolved in 25 mL of methanol and 0.58 g, 2.0 mmol of Ni(NO3)2·6H2O were added to it. The resultant mixture was stirred for 15 min at 50 °C to obtain a dark green solution. A clear solution was obtained by filtering the mixture and the solvent was then allowed to evaporate slowly. After 1 d, green X-ray quality single crystals were isolated. The isolated compound was then washed 3 times with cold methanol and dried in open air.
Yield: 0.394 g (74%, with respect to Ni(II)). Anal. Calcd for (2) C36H52N6Ni3O20: C, 40.60; H, 4.92; N, 7.89. Found: C, 40.54; H, 4.88; N, 7.85. IR (KBr; cm1): 3386 ν(OH), 2964 ν(NH), 1603 ν(C=O), 1387 ν(NO3) and 1068 ν( N–N). ESI-MS(-): m/z 939 [2-2(NO3)-H] (100%).

3.3. X-ray Measurements

Single crystals of complexes 1A and 2 appropriate for X-ray diffraction analysis were immersed in cryo-oil, mounted in Nylon loops, and measured at 296 K. A Bruker AXS PHOTON 100 diffractometer with graphite monochromated Mo-Kα (λ 0.71073) radiation was used to collect the intensity data. Data collections were recorded using omega scans of 0.5° per frame and full sphere of data were obtained. Cell parameters were retrieved using Bruker SMART [49] software and the data were refined using Bruker SAINT [49] on all the observed reflections. Absorption corrections were done using SADABS [49]. Structures were solved by direct methods by applying SIR97 [50] and refined with SHELXL2014 [51]. Calculations were carried out using WinGX v2014.1 [52]. All non-hydrogen atoms were refined anisotropically. Those H-atoms bonded to carbon were placed in the model at geometrically calculated positions and refined using a riding model. Uiso(H) were defined as 1.2Ueq of the parent carbon atoms for phenyl and methyne residues and 1.5Ueq of the parent carbon atoms for the methyl groups. Least square refinements were employed with anisotropic thermal motion parameters for all the non-hydrogen atoms and isotropic for the remaining atoms.

3.4. Catalytic studies

All catalytic tests were run under atmospheric ambiance under with the following conditions for each essay: 1.0–5.0 mol% (0.1–0.5 μmol) of the catalyst precursor 1 or 2 (usually 1 mol%) under solvent-free condition contained 2 mmol of nitroethane and 1 mmol of aldehyde, in that order, or added solvent (2 mL) for reaction in a particular solvent. The reaction mixture was stirred at the particular temperature for the required duration. The solvent was then evaporated, the residue was dissolved in CDCl3, and analyzed by 1H NMR. In the case of solvent-free conditions, the reaction mixture was dissolved in CDCl3 and analyzed by 1H NMR. A previously reported method was employed to determine the yield of the β-nitroalkanol product (relatively to the aldehyde) by 1H NMR [47,48]. To verify the adequacy of the procedure, a number of 1H NMR analyses were performed in the presence of 1,2-dimethoxyethane as an internal standard, added to the CDCl3 solution, giving yields similar to those obtained by the above method. Moreover, the internal standard method also confirmed the absence of other side products formed. The ratio between the syn and anti isomers was also determined by 1H NMR spectroscopy. The values of vicinal coupling constants (for the β-nitroalkanol products), in the 1H NMR spectra, between the α-N–C–H, and the α-O–C–H protons identify the isomers, being J = 7–9 or 3.2–4 Hz for the syn or anti isomers, respectively [53,54].

4. Conclusions

Binuclear and trinuclear aroylhydrazone Ni(II) complexes in two different tautomeric forms (enol and keto) [Ni2(L1)2(MeOH)4] (1) and [Ni3(HL2)2(CH3OH)8]·(NO3)2 (2), where H2L1 = 2-hydroxy(2-hydroxybenzylidene)benzohydrazide, and H3L2 = (2,3-dihydroxybenzylidene)-2-hydroxybenzohydrazide, have been synthesized and fully characterized. They act as efficient catalysts for the solvent-free Henry reaction of β-nitroalkanols formation with excellent yields and diastereoselectivity. The avoidance of a solvent, either organic or ionic liquid, in this reaction brings a number of benefits render it a safer, more economical, and environmentally benign C–C coupling process.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/9/6/554/s1, Figure S1: 1H NMR spectrum for the nitroalol product of nitroethane and benzaldehyde in CDCl3 (Table 3, entry 10); CCDC 1914037-1914038 for 1A and 2 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif.

Author Contributions

Conceptualization, M.S.; data curation, T.R.B.; formal analysis, T.R.B.; funding acquisition, A.J.L.P. and L.M.D.R.S.M.; investigation, M.S.; methodology, M.S. and T.R.B.; project administration, A.J.L.P. and L.M.D.R.S.M.; resources, A.J.L.P.; software, M.S.; supervision, M.S. and L.M.D.R.S.M.; visualization, L.M.D.R.S.M.; writing—original draft, M.S. and T.R.B.; writing—review and editing, A.J.L.P. and L.M.D.R.S.M.

Funding

Authors are grateful to the Fundação para a Ciência e a Tecnologia: FCT (projects UID/QUI/00100/2019, PTDC/QEQ-ERQ/1648/2014 and PTDC/QEQ-QIN/3967/2014), Portugal, for financial support.

Acknowledgments

M.S. acknowledges the FCT and IST for a working contract “DL/57/2017” (Contract no. IST-ID/102/2018). Authors are thankful to the Portuguese NMR Network (IST-UL Centre) for access to the NMR facility and the IST Node of the Portuguese Network of mass-spectrometry for the ESI-MS measurements.

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. Christensen, C.; Juhl, K.; Jorgensen, K.A. Catalytic Asymmetric Henry Reactions—A Simple Approach to Optically Active β-Nitro α-Hydroxy Esters. Chem. Commun. 2001, 2222–2223. [Google Scholar] [CrossRef]
  2. Ballini, R.; Palmieria, A.; Righi, P. Highly Efficient One- or Two-Step Sequences for the Synthesis of Fine Chemicals from Versatile Nitroalkanes. Tetrahedron. 2007, 63, 12099–12121. [Google Scholar] [CrossRef]
  3. Ballini, R.; Barboni, L.; Giarlo, G. Nitroalkanes in Aqueous Medium as an Efficient and Eco-Friendly Source for the One-Pot Synthesis of 1,4-Diketones, 1,4-Diols, δ-Nitroalkanols, and Hydroxytetrahydrofurans. J. Org. Chem. 2003, 68, 9173–9176. [Google Scholar] [CrossRef] [PubMed]
  4. Naïli, H.; Hajlaoui, F.; Mhiri, T.; Mac Leod, T.C.O.; Kopylovich, M.N.; Mahmudov, K.T.; Pombeiro, A.J.L. 2-Dihydromethylpiperazinediium-MII (MII = CuII, FeII, CoII, ZnII) Double Sulfates and their Catalytic Activity in Diastereoselective Nitroaldol (Henry) Reaction. Dalton Trans. 2013, 42, 399–406. [Google Scholar] [CrossRef] [PubMed]
  5. Dhakshinamoorthy, A.; Opanasenko, M.; Cejka, J.; Garcia, H. Metal Organic Frameworks as Solid Catalysts in Condensation Reactions of Carbonyl Groups. Adv. Synth. Catal. 2013, 355, 247–268. [Google Scholar] [CrossRef]
  6. Doyle, A.G.; Jacobsen, E.N. Small-Molecule H-Bond Donors in Asymmetric Catalysis. Chem. Rev. 2007, 107, 5713–5743. [Google Scholar] [CrossRef] [PubMed]
  7. Kopylovich, M.N.; Mac Leod, T.C.O.; Mahmudov, K.T.; Guedes da Silva, M.F.C.; Pombeiro, A.J.L. Zinc(II) ortho-Hydroxyphenylhydrazo-β-diketonate Complexes and their Catalytic Ability Towards Diastereoselective Nitroaldol (Henry) Reaction. Dalton Trans. 2011, 40, 5352–5361. [Google Scholar] [CrossRef] [PubMed]
  8. Shibasaki, M.; Kanai, M.; Matsunaga, S.; Kumagai, N. Multimetallic Multifunctional Catalysts for Asymmetric Reactions: In Bifunctional Molecular Catalysis; Ikariya, T., Shibasaki, M., Eds.; Topics in Organometallic Chemistry; Springer: Berlin, Germany, 2011; Volume 37. [Google Scholar]
  9. Kopylovich, M.N.; Mizar, A.; Guedes da Silva, M.F.C.; Mac Leod, T.C.O.; Mahmudov, K.T.; Pombeiro, A.J.L. Template Syntheses of Copper(II) Complexes from Arylhydrazones of Malononitrile and their Catalytic Activity towards Alcohol Oxidations and the Nitroaldol Reaction: Hydrogen Bond-Assisted Ligand Liberation and E/Z Isomerisation. Chem. Eur. J. 2013, 19, 588–600. [Google Scholar] [CrossRef]
  10. Xu, K.; Lai, G.; Zha, Z.; Pan, S.; Chen, H.; Wang, Z. A Highly anti-Selective Asymmetric Henry Reaction Catalyzed by a Chiral Copper Complex: Applications to the Syntheses of (+)-Spisulosine and a Pyrroloisoquinoline Derivative. Chem. Eur. J. 2012, 18, 12357–12362. [Google Scholar] [CrossRef]
  11. Yao, L.; Wei, Y.; Wang, P.; He, W.; Zhang, S. Promotion of Henry Reactions Using Cu(OTf)2 and a Sterically Hindered Schiff base: Access to Enantioenriched β-Hydroxynitroalkanes. Tetrahedron 2012, 68, 9119–9124. [Google Scholar] [CrossRef]
  12. Ono, N. The Nitro Group in Organic Synthesis Willey-VCH; John Wiley & Sons, Inc.: New York, NY, USA, 2001. [Google Scholar]
  13. Yamaguchi, M.; Shiraishi, T.; Hirama, M. Asymmetric Michael Addition of Malonate Anions to Prochiral Acceptors Catalyzed by l-Proline Rubidium Salt. J. Org. Chem. 1996, 61, 3520–3530. [Google Scholar] [CrossRef]
  14. Varma, R.S.; Dahiya, R.; Kumar, S. Microwave-assisted Henry reaction: Solventless synthesis of conjugated nitroalkenes. Tetrahedron Lett. 1997, 38, 5131–5134. [Google Scholar] [CrossRef]
  15. Noland, W.E. The NEF Reaction. Chem. Rev. 1955, 55, 137–155. [Google Scholar] [CrossRef]
  16. Sasai, H.; Suzuki, T.; Arai, S.; Arai, T.; Shibasaki, M. Basic character of rare earth metal alkoxides. Utilization in catalytic carbon-carbon bond-forming reactions and catalytic asymmetric nitroaldol reactions. J. Am. Chem. Soc. 1992, 114, 4418–4420. [Google Scholar] [CrossRef]
  17. Das, A.; Kureshy, R.I.; Prathap, K.J.; Choudhary, M.K.; Rao, G.V.; Khan, N.-U.H.; Abdi, S.H.; Bajaj, H.C. Chiral recyclable Cu(II)-catalysts in nitroaldol reaction of aldehydes with various nitroalkanes and its application in the synthesis of a valuable drug (R)-isoproterenol. Appl. Catal. A: Gen. 2013, 459, 97–105. [Google Scholar] [CrossRef]
  18. Das, A.; Kureshy, R.I.; Subramanian, P.S.; Khan, N.-U.H.; Abdi, S.H.R.; Bajaj, H.C. Synthesis and characterization of chiral recyclable dimeric copper(ii)–salen complexes and their catalytic application in asymmetric nitroaldol (Henry) reaction. Catal. Sci. Technol. 2014, 4, 411–418. [Google Scholar] [CrossRef]
  19. Selvakumar, P.M.; Suresh, E.; Subramanian, P. Single stranded helical supramolecular architecture with a left handed helical water chain in ternary copper(II) tryptophan/diamine complexes. Polyhedron 2009, 28, 245–252. [Google Scholar] [CrossRef]
  20. Kitagaki, S.; Ueda, T.; Mukai, C. Planar chiral [2.2]paracyclophane-based bis(thiourea) catalyst: Application to asymmetric Henry reaction. Chem. Commun. 2013, 49, 4030. [Google Scholar] [CrossRef]
  21. Nitabaru, T.; Nojiri, A.; Kobayashi, M.; Kumagai, N.; Shibasaki, M. anti-Selective Catalytic Asymmetric Nitroaldol Reaction via a Heterobimetallic Heterogeneous Catalyst. J. Am. Chem. Soc. 2009, 131, 13860–13869. [Google Scholar] [CrossRef]
  22. Karmakar, A.; Da Silva, M.F.C.G.; Pombeiro, A.J.L. Zinc metal–organic frameworks: Efficient catalysts for the diastereoselective Henry reaction and transesterification. Dalton Trans. 2014, 43, 7795–7810. [Google Scholar] [CrossRef]
  23. Kehat, T.; Portnoy, M. Polymer-Supported Proline-Decorated Dendrons: Dendritic Effect in Asymmetric Aldol Reaction. Chem. Commun. 2007, 2823–2825. [Google Scholar] [CrossRef] [PubMed]
  24. McNulty, J.; Steere, J.A.; Wolf, S. The ultrasound promoted Knoevenagel condensation of aromatic aldehydes. Tetrahedron Lett. 1998, 39, 8013–8016. [Google Scholar] [CrossRef]
  25. Neelakandeswari, N.; Sangami, G.; Emayavaramban, P.; Karvembu, R.; Dharmaraj, N.; Kim, H.Y. Mesoporous nickel hydroxyapatite nanocomposite for microwave-assisted Henry reaction. Tetrahedron Lett. 2012, 53, 2980–2984. [Google Scholar] [CrossRef]
  26. Hazra, S.; Karmakar, A.; Silva, M.D.F.C.G.D.; Dlháň, L.; Boča, R.; Pombeiro, A.J.L. Sulfonated Schiff base dinuclear and polymeric copper(ii) complexes: Crystal structures, magnetic properties and catalytic application in Henry reaction. New J. Chem. 2015, 39, 3424–3434. [Google Scholar] [CrossRef]
  27. Gupta, M.; De, D.; Pal, S.; Pal, T.K.; Tomar, K. A porous two-dimensional Zn(ii)-coordination polymer exhibiting SC–SC transmetalation with Cu(ii): Efficient heterogeneous catalysis for the Henry reaction and detection of nitro explosives. Dalton Trans. 2017, 46, 7619–7627. [Google Scholar] [CrossRef] [PubMed]
  28. Martins, N.M.; Mahmudov, K.T.; Da Silva, M.F.C.G.; Martins, L.M.; Guseinov, F.I.; Pombeiro, A.J. 1D Zn(II) coordination polymer of arylhydrazone of 5,5-dimethylcyclohexane-1,3-dione as a pre-catalyst for the Henry reaction. Catal. Commun. 2016, 87, 49–52. [Google Scholar] [CrossRef]
  29. Tanaka, K.; Toda, F. Solvent-Free Organic Synthesis. Chem. Rev. 2000, 100, 1025–1074. [Google Scholar] [CrossRef]
  30. Tanaka, K.; Hachiken, S. Enantioselective Henry reaction catalyzed by trianglamine–Cu(OAc)2 complex under solvent-free conditions. Tetrahedron Lett. 2008, 49, 2533–2536. [Google Scholar] [CrossRef]
  31. Angelini, T.; Ballerini, E.; Bonollo, S.; Curini, M.; Lanari, D. A new sustainable protocol for the synthesis of nitroaldol derivatives via Henry reaction under solvent-free conditions. Green Chem. Lett. Rev. 2014, 7, 11–17. [Google Scholar] [CrossRef]
  32. Huseynov, F.E.; Shamilov, N.T.; Voronina, A.A.; Buslaeva, T.M.; Mahmudov, K.T.; Da Silva, M.F.C.G.; Sutradhar, M.; Kopylovich, M.N.; Pombeiro, A.J.L. Lanthanide derivatives comprising arylhydrazones of β-diketones: Cooperative E/Z isomerization and catalytic activity in nitroaldol reaction. Dalton Trans. 2015, 44, 5602–5610. [Google Scholar]
  33. Rocha, B.G.M.; Mac Leod, T.C.O.; Guedes da Silva, M.F.C.; Luzyanin, K.V.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. NiII, CuII and ZnII Complexes with a Sterically Hindered Scorpionate Ligand (TpmsPh) and Catalytic Application in the Diasteroselective Nitroaldol (Henry) Reaction. Dalton Trans. 2014, 43, 15192–15200. [Google Scholar] [CrossRef] [PubMed]
  34. Pettinari, C.; Marchetti, F.; Cerquetella, A.; Pettinari, R.; Monari, M.; Mac Leod, T.C.O.; Martins, L.M.D.R.S.; Pombeiro, A.J.L. Coordination chemistry of the (h6-cymene)ruthenium(II) fragment with bis-, tris-, and tetrakis(pyrazol-1-yl)borate ligands: Synthesis, structural, electrochemical and catalytic diastereoselective nitroaldol reaction studies. Organometallics 2011, 30, 1616–1626. [Google Scholar] [CrossRef]
  35. Ribeiro, A.P.C.; Karabach, Y.Y.; Martins, L.M.D.R.S.; Mahmoud, A.G.; Guedes da Silva, M.F.C.; Pombeiro, A.J.L. Nickel(II)-2-amino-4-alkoxy-1,3,5-triazapentadienato complexes as catalysts for Heck and Henry reactions. RSC Adv. 2016, 6, 29159–29163. [Google Scholar] [CrossRef]
  36. Sutradhar, M.; Da Silva, M.F.C.G.; Pombeiro, A.J. A new cyclic binuclear Ni(II) complex as a catalyst towards nitroaldol (Henry) reaction. Catal. Commun. 2014, 57, 103–106. [Google Scholar] [CrossRef]
  37. Arunachalam, R.; Aswathi, C.S.; Das, A.; Kureshy, R.I.; Subramanian, P.S. ChemInform Abstract: Diastereoselective Nitroaldol Reaction Catalyzed by Binuclear Copper(II) Complexes in Aqueous Medium. ChemPlusChem 2015, 80, 209–216. [Google Scholar] [CrossRef]
  38. Karmakar, A.; Da Silva, M.F.C.G.; Hazra, S.; Pombeiro, A.J.L. Zinc amidoisophthalate complexes and their catalytic application in the diastereoselective Henry reaction. New J. Chem. 2015, 39, 3004–3014. [Google Scholar] [CrossRef]
  39. Sutradhar, M.; Martins, L.M.D.R.S.; Da Silva, M.F.C.G.; Mahmudov, K.T.; Liu, C.-M.; Pombeiro, A.J.L. Trinuclear Cu II Structural Isomers: Coordination, Magnetism, Electrochemistry and Catalytic Activity towards the Oxidation of Alkanes. Eur. J. Inorg. Chem. 2015, 2015, 3959–3969. [Google Scholar] [CrossRef]
  40. Sutradhar, M.; Martins, L.M.D.R.S.; Guedes da Silva, M.F.C.; Pombeiro, A.J.L. Vanadium Complexes: Recent Progress in Oxidation Catalysis. Coord. Chem. Rev. 2015, 301–302, 200–239. [Google Scholar] [CrossRef]
  41. Sutradhar, M.; Pombeiro, A.J.L. Coordination Chemistry of Non-oxido, Oxido and Dioxidovanadium(IV/V) Complexes with Azine Fragment Ligands. Coord. Chem. Rev. 2014, 265, 89–124. [Google Scholar] [CrossRef]
  42. Sutradhar, M.; Barman, T.R.; Ghosh, S.; Drew, M.G. Synthesis of a mononuclear oxidovanadium(V) complex by bridge-splitting of the corresponding binuclear precursor. J. Mol. Struct. 2012, 1020, 148–152. [Google Scholar] [CrossRef]
  43. Sutradhar, M.; Alegria, E.C.B.A.; Mahmudov, K.T.; Da Silva, M.F.C.G.; Pombeiro, A.J.L. Iron(iii) and cobalt(iii) complexes with both tautomeric (keto and enol) forms of aroylhydrazone ligands: Catalysts for the microwave assisted oxidation of alcohols. RSC Adv. 2016, 6, 8079–8088. [Google Scholar] [CrossRef]
  44. Sutradhar, M.; Barman, T.R.; Mukherjee, G.; Drew, M.G.; Ghosh, S. Synthesis, structural characterization and electrochemical activity of oxidovanadium(IV/V) complexes of a diprotic ONS chelating ligand. Inorganica Chim. Acta 2010, 363, 3376–3383. [Google Scholar] [CrossRef]
  45. Sutradhar, M.; Carrella, L.M.; Rentschler, E. A Discrete μ4-Oxido Tetranuclear Iron(III) Cluster. Eur. J. Inorg. Chem. 2012, 2012, 4273–4278. [Google Scholar] [CrossRef]
  46. Kopylovich, M.N.; Mahmudov, K.T.; Silva, M.F.C.G.; Martins, L.M.D.R.S.; Kuznetsov, M.L.; Silva, T.F.S.; Fraústo da Silva, J.J.R.; Pombeiro, A.J.L. Trends in properties of para-substituted 3-(phenylhydrazo)pentane-2,4-diones. J. Phys. Org. Chem. 2011, 24, 764–773. [Google Scholar] [CrossRef]
  47. Sutradhar, M.; Martins, L.M.; Da Silva, M.F.C.G.; Pombeiro, A.J. Oxidovanadium complexes with tridentate aroylhydrazone as catalyst precursors for solvent-free microwave-assisted oxidation of alcohols. Appl. Catal. A Gen. 2015, 493, 50–57. [Google Scholar] [CrossRef]
  48. Sutradhar, M.; Martins, L.; Da Silva, M.F.C.G.; Alegria, E.C.B.A.; Liu, C.-M.; Pombeiro, A.J.L. Dinuclear Mn(ii,ii) complexes: Magnetic properties and microwave assisted oxidation of alcohols. Dalton Trans. 2014, 43, 3966. [Google Scholar] [CrossRef]
  49. Bruker, APEX2 & SAINT; AXS Inc.: Madison, WI, USA, 2004.
  50. Altomare, A.; Camalli, M.; Giacovazzo, C.; Moliterni, A.G.; Polidori, G.; Spagna, R.; Moliterni, A.G.G.; Burla, M.C.; Cascarano, G.L.; Guagliardi, A.; et al. SIR 97: A new tool for crystal structure determination and refinement. J. Appl. Crystallogr. 1999, 32, 115–119. [Google Scholar] [CrossRef]
  51. Sheldrick, G.M. A Short History of SHELX. Acta Crystallogr. Sect. A. 2008, 64, 112–122. [Google Scholar] [CrossRef]
  52. Farrugia, L.J. WinGX and ORTEP for Windows: An update. J. Appl. Crystallogr. 2012, 45, 849–854. [Google Scholar] [CrossRef]
  53. Cwik, A.; Fuchs, A.; Hell, Z.; Clacens, J.-M.; Clacens, J. Nitroaldol-Reaction of Aldehydes in the Presence of Non-Activated Mg:Al 2:1 Hydrotalcite: A Possible New Mechanism for the Formation of 2-Aryl-1,3-dinitropropanes. Tetrahedron 2005, 36, 4015–4021. [Google Scholar] [CrossRef]
  54. Bulbule, V.J.; Deshpande, V.H.; Velu, S.; Sudalai, A.; Sivasankar, S.; Sathe, V. Heterogeneous Henry reaction of aldehydes: Diastereoselective synthesis of nitroalcohol derivatives over Mg-Al hydrotalcites. Tetrahedron 1999, 55, 9325–9332. [Google Scholar] [CrossRef]
Scheme 1. Syntheses of 1 and 2.
Scheme 1. Syntheses of 1 and 2.
Catalysts 09 00554 sch001
Figure 1. Molecular structure of compound 1A with atom numbering scheme. (a) nickel(II) coordinated with two DMF, (b) nickel(II) coordinated with one DMF and one water molecule. Sky-blue doted lines represent hydrogen bonds. Symmetry codes (i) −x, 1 − y, 1 − z (ii) 1 − x, 2 − y, 2 − z.
Figure 1. Molecular structure of compound 1A with atom numbering scheme. (a) nickel(II) coordinated with two DMF, (b) nickel(II) coordinated with one DMF and one water molecule. Sky-blue doted lines represent hydrogen bonds. Symmetry codes (i) −x, 1 − y, 1 − z (ii) 1 − x, 2 − y, 2 − z.
Catalysts 09 00554 g001
Figure 2. Molecular structure of compound 2 with atom numbering scheme. Sky-blue doted lines represent hydrogen bonds. Symmetry code (i) 1 − x, 1 − y, −z.
Figure 2. Molecular structure of compound 2 with atom numbering scheme. Sky-blue doted lines represent hydrogen bonds. Symmetry code (i) 1 − x, 1 − y, −z.
Catalysts 09 00554 g002
Scheme 2. Nitroaldol condensation reaction of nitroethane and benzaldehydes.
Scheme 2. Nitroaldol condensation reaction of nitroethane and benzaldehydes.
Catalysts 09 00554 sch002
Scheme 3. Nitroaldol condensation reaction of various aldehydes and different nitroalkanes.
Scheme 3. Nitroaldol condensation reaction of various aldehydes and different nitroalkanes.
Catalysts 09 00554 sch003
Scheme 4. Proposed catalytic cycle for the β-nitroalkanol formation in the Ni(II) catalyzed nitroaldol reaction.
Scheme 4. Proposed catalytic cycle for the β-nitroalkanol formation in the Ni(II) catalyzed nitroaldol reaction.
Catalysts 09 00554 sch004
Table 1. Crystal data and structure refinement details for complexes 1A and 2.
Table 1. Crystal data and structure refinement details for complexes 1A and 2.
1A2
Empirical formulaC20H24N4NiO5.50C18H26N3Ni1.50O10
Formula Weight 467.14532.48
Crystal system TriclinicTriclinic
Space groupP¯1P¯1
Temperature/K 296 (2)296 (2)
a12.766 (3)9.114 (3)
b13.714 (4)10.821 (3)
c14.344 (4)13.221 (4)
α/° 84.87 (1)71.721 (10)
β63.594 (10)83.309 (11)
γ/° 79.445 (9)71.328 (10)
V3) 2211.2 (10) 1172.8 (6)
Z42
Dcalc (g cm−3) 1.4031.508
μ(Mo Kα) (mm−1)0.921.27
Rfls. collected/unique/observed35885/8156/668310924/4209/2646
Rint0.0290.091
Final R1 a, wR2 b (I ≥ 2σ)0.038, 0.1080.093, 0.224
Goodness-of-fit on F21.031.05
a R = Σ||Fo| − |Fc||/Σ|Fo|; b wR(F2) = [Σw(|Fo|2 − |Fc|2)2/Σw|Fo|4]½, w = 1/[ σ2(Fo2) + (aP)2 + bP].
Table 2. Selected bond distances (Å) and angles (°) in complexes 1A and 2.
Table 2. Selected bond distances (Å) and angles (°) in complexes 1A and 2.
1A
Ni1—N11.995 (2)N1—Ni1—O189.97 (8)N5—Ni2—O689.33 (8)
Ni1—O12.0231 (17)N1—Ni1—O1i170.34 (8)N5—Ni2—O6ii169.38 (8)
Ni1—O1i2.0256 (16)O1—Ni1—O1i80.43 (7)O6—Ni2—O6ii80.07 (8)
Ni1—O22.039 (2)N1—Ni1—O278.97 (9)N5—Ni2—O779.06 (8)
Ni1—O42.128 (2)O1—Ni1—O2168.93 (7)O6—Ni2—O7168.18 (8)
Ni1—O52.1753 (18)O1i—Ni1—O2110.62 (7)O6ii—Ni2—O7111.52 (8)
Ni2—N52.009 (2)N1—Ni1—O495.68 (9)N5—Ni2—O992.27 (8)
Ni2—O62.0342 (18)O1—Ni1—O489.38 (8)O6—Ni2—O993.83 (8)
Ni2—O6ii2.0488 (18)O1i—Ni1—O485.34 (8)O6ii—Ni2—O988.04 (8)
Ni2—O72.059 (2)O2—Ni1—O492.13 (8)O7—Ni2—O984.51 (8)
Ni2—O92.0975 (19)N1—Ni1—O591.79 (8)N5—Ni2—O1092.08 (8)
Ni2—O102.1161 (19)O1—Ni1—O590.17 (7)O6—Ni2—O1094.69 (8)
N1—N21.385 (3)O1i—Ni1—O587.23 (7)O6ii—Ni2—O1089.24 (8)
N5—N61.397 (3)O2—Ni1—O589.74 (8)O7—Ni2—O1088.00 (8)
O2—C81.276 (3)O4—Ni1—O5172.52 (7)O9—Ni2—O10170.47 (9)
O7—C281.267 (3)Ni1—O1—Ni1i99.57 (7)Ni2—O6—Ni2ii99.93 (8)
2
Ni1—O42.089 (7)N2—Ni1—O5170.5 (2)O3—Ni2—O3iii180.0
Ni1—N21.991 (6)N2—Ni1—O290.7 (2)O3—Ni2—O2iii99.4 (2)
Ni1—O52.025 (6)O5—Ni1—O298.7 (2)O3iii—Ni2—O2iii80.6 (2)
Ni1—O22.031 (5)N2—Ni1—O178.7 (2)O3—Ni2—O280.6 (2)
Ni1—O12.088 (6)O5—Ni1—O191.9 (2)O3iii—Ni2—O299.4 (2)
Ni1—O62.092 (6)O2—Ni1—O1169.4 (2)O2iii—Ni2—O2180.0
Ni2—O31.948 (6)N2—Ni1—O487.6 (3)O3—Ni2—O788.8 (3)
Ni2—O3iii1.948 (5)O5—Ni1—O493.5 (3)O3iii—Ni2—O791.2 (3)
Ni2—O2iii2.108 (5)O2—Ni1—O493.1 (2)O2iii—Ni2—O793.3 (2)
Ni2—O22.108 (5)O1—Ni1—O486.8 (2)O2—Ni2—O786.7 (2)
Ni2—O72.143 (8)N2—Ni1—O691.9 (3)O3—Ni2—O7iii91.2 (3)
Ni2—O7iii2.143 (8)O5—Ni1—O686.7 (3)O3iii—Ni2—O7iii88.8 (3)
N1—N21.387 (8)O2—Ni1—O688.8 (2)O2iii—Ni2—O7iii86.7 (2)
C1—O11.250 (9)O1—Ni1—O691.3 (3)O2—Ni2—O7iii93.3 (2)
--O4—Ni1—O6178.0 (3)O7—Ni2—O7iii180.0
--Ni1—O2—Ni2125.2 (2)--
Symmetry codes: (i) −x, −y + 1, −z + 1; (ii) −x + 1, −y + 2, −z + 2; (iii) −x + 1, −y + 1, −z.
Table 3. Catalytic activity of 1 and 2 in the model nitroaldol condensation reaction a of nitroethane and benzaldehyde.
Table 3. Catalytic activity of 1 and 2 in the model nitroaldol condensation reaction a of nitroethane and benzaldehyde.
Entry.CatalystTime
(h)
Amount of Catalyst
(mol%)
Temp. (°C)SolventYield or Conversion (%) bSelectivity c
syn: anti
TOF/h−1 d
11241.0ambient-86.670:303.6
21241.0ambientH2O80.270:303.3
31241.0ambientMeOH72.264:363.0
41241.0ambientMeCN56.458:422.4
52241.0ambient-89.272:283.7
62241.0ambientH2O82.470:303.4
72241.0ambientMeOH72.666:343.0
82241.0ambientMeCN58.060:402.4
92241.040-90.872:283.8
102241.060-94.077:233.9
112241.075-90.676:243.8
12Blank2460----
13Ni(OAc)2241.060-962:380.3
142242.060-94.276:241.9
152243.060-92.874:261.3
162245.060-91.671:290.8
17261.060-44.468:327.4
182121.060-60.872:285.1
192481.060-92.674:261.9
a Experimental conditions: Benzaldehyde (1 mmol); nitroethane (2 mmol); solvent (2 mL); b Determined using 1H NMR analysis; c Molar ratios, calculated using 1H NMR (see Figure S1 for entry 10); d Estimated as moles of syn- and anti-β-nitroethanol/mol of 1 or 2, per hour.
Table 4. Solvent-free nitroaldol condensation of different aldehydes and nitroethane catalyzed by 2 a.
Table 4. Solvent-free nitroaldol condensation of different aldehydes and nitroethane catalyzed by 2 a.
EntrySubstrateYield or Conversion (%) bsyn: anti ratio cTOF/h−1 d
1 Catalysts 09 00554 i00194.077:233.9
2 Catalysts 09 00554 i00260.872:282.5
3 Catalysts 09 00554 i00356.470:302.3
4 Catalysts 09 00554 i00474.076:243.1
5 Catalysts 09 00554 i00588.4 74:263.7
6 Catalysts 09 00554 i00670.2 70:302.9
7 Catalysts 09 00554 i00797.080:204.0
8CH3CHO66.672:282.8
9CH3CH2CHO78.274:263.2
a Experimental conditions: Catalyst 2, 1.0 mol%, aldehyde (1 mmol), nitroethane (2 mmol), 24 h, 60 °C; b Determined using 1H NMR analysis (see Experimental); c Molar ratio, calculated using 1H NMR; d Estimated as moles of syn- and anti-β-nitroethanol/mol of 2, per hour.
Table 5. Solvent-free nitroaldol condensation of different aldehydes and nitroalkanes catalyzed by 2 a.
Table 5. Solvent-free nitroaldol condensation of different aldehydes and nitroalkanes catalyzed by 2 a.
EntryAldehydeNitroalkaneYield or Conversion /% bsyn: anti ratio cTOF/h−1 d
1 Catalysts 09 00554 i008nitromethane94.2-3.9
2nitroethane94.077:233.9
3nitropropane54.268:322.5
4 Catalysts 09 00554 i009nitromethane56.6-2.3
5nitroethane44.466:341.8
6nitropropane38.4 58:421.6
a Experimental conditions: Catalyst 2, (1.0 mol%), aldehyde (1 mmol), nitroalkane (2 mmol), 24 h, 60 °C; b Determined using 1H NMR analysis (see Experimental); c Molar ratio, determined using 1H NMR; d Estimated as moles of syn- and anti-β-nitroethanol/mol of 2, per hour.

Share and Cite

MDPI and ACS Style

Sutradhar, M.; Roy Barman, T.; Pombeiro, A.J.L.; Martins, L.M.D.R.S. Ni(II)-Aroylhydrazone Complexes as Catalyst Precursors Towards Efficient Solvent-Free Nitroaldol Condensation Reaction. Catalysts 2019, 9, 554. https://doi.org/10.3390/catal9060554

AMA Style

Sutradhar M, Roy Barman T, Pombeiro AJL, Martins LMDRS. Ni(II)-Aroylhydrazone Complexes as Catalyst Precursors Towards Efficient Solvent-Free Nitroaldol Condensation Reaction. Catalysts. 2019; 9(6):554. https://doi.org/10.3390/catal9060554

Chicago/Turabian Style

Sutradhar, Manas, Tannistha Roy Barman, Armando J. L. Pombeiro, and Luísa M.D.R.S. Martins. 2019. "Ni(II)-Aroylhydrazone Complexes as Catalyst Precursors Towards Efficient Solvent-Free Nitroaldol Condensation Reaction" Catalysts 9, no. 6: 554. https://doi.org/10.3390/catal9060554

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop