Sie sind auf Seite 1von 17

CONTROLLING THE ACTIVITY AND CHEMOSELECTIVITY IN THE CINNAMALDEHYDE HYDROGENATION BY INSERTION OF NONNOBLE METALS IN THE MATRIX OF HYDROALCITE-LIKE MATERIALS

Alexandru Chirieac, Brindusa Dragoi, Adrian Ungureanu, Alina Moscu Corcodel, Constantin Rudolf, Alexandra Sasu, Emil Dumitriu1 Department of Organic and Biochemical Engineering, Faculty of Chemical Engineering and Environmental Protection, Gheorghe Asachi Technical University of Iasi, 73 Prof.dr.docent Dimitrie Mangeron Street,700050-Iasi, Romania,

Abstract

The purpose of this study was to prepare new nanocomposite catalysts of Metal/LDHs type by mild reduction of transition metals Co2+ and Cu2+ cations substituted in the matrix of ZnAl-LDH at 150C and atmospheric pressure, under hydrogen flow. To this end, a series of ZnCuCoAl layered double hydroxides (LDH) precursors were synthesized by co-precipitation method. M2+/M3+ (i.e., Zn2+/Al3+) molar ratios were kept constant while the Cu2+/Co2+ molar ratios were varied in order to investigate their influence on the structural and textural properties of the resulted catalysts. The LDH precursors were systematically characterized by different techniques such as powder XRD, nitrogen physisorption, FT-IR and DR UV-Vis spectroscopy as well as temperature programmed reduction (TPR). The characterization results indicate that all the precursors present well-crystallized layered structures and a high degree of substitution of cobalt and copper cations in the ZnAl matrix. Upon the mild reduction treatment the layered structure of precursors was maintained, whereas a certain part of transition cations was reduced to metallic centers acting as highly dispersed catalytic active sites. The effect of chemical composition of the ZnCuCoAl LDHs on the catalytic properties in the selective hydrogenation of cinnamaldehyde was studied in propylene carbonate at 150C and atmospheric pressure. It was observed that catalytic activity gradually increases with the Cu content of LDH, whereas the selectivity to unsaturated alcohol does not depend on the chemical composition. A substantial high chemoselectivity of LDHs can be however attained by replacing Cu with Ni in ZnCoAl LDH precursors at the expense of decreased activity.. Keywords: cinnamaldehyde hydrogenation, layered double hydroxides, Co-based catalysts, Cubased catalysts
1

Author to whom all correspondence should be addressed: E-mail: edumitri@tuiasi.ro; Phone: +40 232278683; Fax: +40 232271311

1. Introduction

The intensive demand for new environmentally friendly materials and technologies led to the development of the catalysis field which now provides sustainable routes to convert efficiently and economically the raw materials into valuable chemicals, fuels and other products. The synthesis of fine chemicals is an area of great interest where catalysis should be more deeply involved since this industry is under great pressure to develop new efficient processes with low energy and raw materials consumption and minimal wastes. Nonexpensive catalysts based on transitional metals are intensively studied nowadays to meet these requirements. The selective hydrogenation of , -unsaturated aldehydes such as cinnamaldehyde is of great importance in fine chemicals industry especially in flavoring, perfume and pharmaceutical (Castelijins et al. 1998, Tessonnier et al., 2005; Gallezot et al., 1998; Coq et al., 1998; Ponec, 1997) because it leads to the formation of useful products (i.e., cinnamylalcohol (CNOL) and hydrocinnamaldehyde (HCNA)). However, the hydrogenation of C=C bonds is favored from thermodynamic and kinetic considerations. Therefore the hydrogenation of C=O bond to obtain CNOL is a continuous and challenging scientific task since it requires chemoselective catalysts. High selectivity to CNOL was obtained with Ir- and Os-based catalysts (Mki-Arvela et al., 2005), while the noble metals from Pt group display selectivity to CNOL when promoters are used. However, the high cost and low availability of these catalysts cause a serious problem, and therefore enhanced research efforts are made nowadays to develop active, chemoselective and inexpensive catalysts which are based on non-noble metals. In this regard, the synthesis of nanocomposite materials affords new opportunities on the design of more active and selective catalysts. This study is focused on the synthesis of some Me/LDH nanocomposites (Me = Co, Cu) obtained from layered double hydroxide (LDH) precursors with applications in the liquid phase hydrogenation of the cinnamaldehyde. The general formula for the LDHs is [M1x2+Mx3+(OH)2]x+(An)x/nmH2O, where M2+ and M3+ are di- and trivalent cations respectively, including Mg2+, Zn2+, Ni2+, Co2+, or Cu2+ and Al3+, Ga3+, Cr3+ or Fe3+, respectively; x is equal to the molar ratio M2+/(M2++M3+), and An can be simple or complex anions. Generally, the structure of LDHs consists of positively charged layers which are stacked on top of each other (Fig. 1.a). These positive charges are compensated by anions which lie in the interlayer region between two sheets. In the space between layers are also accommodated molecules of crystallization water.

Figure 1. Schematic illustration of Me/LDH nanocomposite forming.

Obviously, the LDHs have been used as calcined materials when a homogeneous mixture of oxide is obtained. It was shown that they possess quite interesting properties such as high stability to thermal treatments and basic and/or redox properties (Cavani et al., 1991; Vaccari, 1998; Vaccari, 1999; Rives, 2001). At the same time, non-calcined LDHs are active catalysts for some base catalytic reactions like as: intramolecular cyclization of cinnamyl derivatives (Finta et al., 2000), alkoxylation (F. Malherbe et al. Appl. Clay Sci. 13 (1998) 451), ester exchange (Y. Tanabe et al., Micropor. Mater. 22 (1998) 399) etc. Also, non-calcined LDHs could be attractive as hydrogenation catalysts since they contain various reducible transition metal cations which can generate catalytically active species well dispersed in the LDH matrix (Dragoi et al., 2010a,b). Redox properties of the metal ions are significantly modified by the matrix nature and the presence co-cations associated with the system and their concentration. Moreover, it was observed that LDHs with transition metals exhibit enhanced catalytic activity associated with the presence of both redox and Lewis sites (Rives et al., 2001). Taking into consideration the above mentioned properties, a series of catalysts based on ZnAl-LDH matrix substituted with transition metal cations Co2+ and Cu2+ was prepared, and then the as-synthesized LDHs are gently reduced at low temperature to form nanocomposite materials consisting of highly dispersed metallic species in crystalline LDH phase (Fig. 1b). The catalytic properties of these nanocomposites were evaluated by test reaction of cinnamaldehyde hydrogenation.

2. Experimental

2.1. Preparation of LDHs

Layered double hydroxides synthesis containing Zn, Cu, Co and Al were prepared using the low supersaturation co-precipitation method at room temperature. Two solutions, first containing metal nitrates ([Zn2+] + [Cu2+] + [Co2+] + [Al3+] = 1.0 M) and a second alkaline solution (NaOH + Na2CO3 = 1.1 M) were simultaneously added in a three-necked round-bottom flask fitted with a pH electrode and containing a small amount of deionized water. The suspension was vigorously stirred and maintained at the desired pH (~ 8) by adjusting the relative flow rates of the two solutions. The final suspension was aged under stirring at room temperature for 12 h. The precipitate was filtered off, and then washed two times with carbonate solution in order to remove nitrate ions from the interlayers, and finally repeatedly washed with sufficient deionized water to remove residual sodium. The resulting precipitates were dried at 40 C for 24 h. The solids are labeled according to the ratio between cooper and cobalt as ZCuCoA XY in which X and Y represents the molar ratio of the substituted elements (Cu and Co) in the ZnAl matrix.

2.2. Characterization of the catalysts

Powder X-ray diffraction (XRD) patterns were recorded on a SEIFERT diffractometer using monochromatized Cu K radiation ( = 1.542 ). Usually, the diffraction data were collected using continuous scan mode with a scan speed of 2 min-1 over the scan range 2 = 2 70 . N2 adsorption and desorption isotherms were obtained at 77 K on Autosorb 1-MP (Quantachrome) instrument. Prior to the measurements, samples (ca. 80 100 mg) were outgassed under high vacuum at 150 C for 3h. The volume of adsorbed N2 was normalized to standard temperature and pressure. FT-IR spectra were recorded on a Perkin-Elmer FT1730 instrument, using KBr pellets; 100 spectra (recorded with a nominal resolution of 4 cm1) were averaged to improve the signal-to-noise ratio. Diffuse reflectance UVvisible (DR UVvis) spectra were recorded on a Shimadzu UV2450 spectrometer equipped with integrating sphere unit (ISR-2200). The spectra were collected at 190 800 nm using BaSO4 as reference. Temperature-programmed reduction (TPR) measurements were carried with a TPDRO 1100 instrument. The reducing agent was H2/Ar (5% vol). The gas flow (50 mlmin1), sample weight (15 20 mg) and heating rate (10 Cmin1) were chosen according to preliminary tests aiming to

optimize resolution of the curves. A 13X molecular sieve trap was used to remove water. Calibration of the instrument was carried out with CuO (Merck).

2.3. Catalytic activity

For each test, the non-calcined catalyst was first reduced at 150 C for 2 h (heating rate of 6 Cmin1) under hydrogen flow (1 L.h1) and then transferred to the reaction system. The catalytic hydrogenation of trans-cinnamaldehyde test was carried out at atmospheric pressure in a three neck glass reactor equipped with reflux condenser and magnetic stirrer (900 rpm) under the following conditions: 1 ml of reagent (7.94 103 mol), 25 ml of propylene carbonate as solvent, and 0.265 g of catalyst, hydrogen flow 1 L.h1, and constant reaction temperature of 150 C. Samples of the reaction mixture were withdrawn periodically and the identification and quantification of reaction products were performed with GC (HP 5890 gas chromatograph, FID detector, DB-5 column) and were based on the retention times and suitable response factors, respectively. In order to confirm the identity of the products, GC-MS analyses were performed with an Agilent 6890N system equipped with an Agilent 5973 MSD detector and a DB-5-ms column.

3. Results and Discussion

3.1. Chemical and physical properties

There are few reports about synthesis and properties of multicomponent LDHs which contain Co in their structure; it could be mentioned some ternay LDHs: CoZnAl (Benito et al,. 2009), CoNiAl (Rives et al., 2003), quaternary LDHs: MgCoNiAl (Tichit et al., 2001), CoCuZnAl (Velu et al, 2001) and even with five elements CuZnCoAlCr (Porta et al., 1995). However, to the best of our knowledge, there are no previous reports about Me/LDHs nanocomposites obtained by mild reduction from layered double hydroxides (LDHs) precursors containing Co2+ and Cu2+ transition metal cations substituted in the matrix of ZnAl-LDH. The characterization of the LDH solids was performed to study the obtained crystalline phases, phase purity, texture as well as to prove the substitution of transitional metals in the brucite-like layers. To this aim, the as-synthesized solids were first characterized from crystallographic point of view. Figure 2 shows the XRD patterns of the LDH precursors with constant Zn/Al ratios, constant M2+/M3+ ratios, but with variable Cu2+/Co2+ ratios.

Figure 2. XRD patterns for the as-made ZnCuCoAl LDHs obtained by co-precipitation

These diffractograms indicate that pure layered double hydroxides were synthesized for all compositions. The shape and symmetry of XRD peaks are characteristic of well-crystallized materials. The XRD patterns show sharp and symmetrical reflections at low 2 angles (planes (003), (006), (110) and (113)) and broad and asymmetric peaks at high 2 angles (planes (012), (015) and (018)). According to literature, these peaks are characteristic for crystalline phases of hydrotalcite-like LDHs (Cavani et al., 1991). The absence of other crystalline impurity phases (e.g., malachite Cu2CO3(OH)2) suggests that ions of Co2+ and Cu2+ are isomorphous substituted in the brucite-like layers. It can be also observed in Figure 2 that ZCuCoA 10 sample (the Co-free sample) exhibits a high crystallinity, which gradually decreases with the increase in the content of Co, in well agreement with Velu et al., 2001. From the XRD diffractograms, the lattice parameters a and c as well as the crystallite sizes were calculated and their values are included Table 1. As it can be observed, with the increase in the copper content the value of parameter a increases, in agreement with the larger ionic radius of Cu2+ with respect to Co2+ in octahedral environments, while the value of parameter c decreases. The molar fractions of Zn2+ and Al3+ are almost constant in all samples, indicating the compliance of Vegards rule. It is worthy of note that upon the mild reduction treatment, the XRD results (not show) show that the hydrotalcite-like structures of LDH precursors were maintained.

Figure 3. Nitrogen adsorption-desorption isotherms for as-made ZnCuCoAl LDHs obtained by co-precipitation

The textural properties such as BET surface area, the average pores diameter and pore volume for the investigated LDH precursors are also shown in Table 1 and are in agreement with those reported in the literature (Benito et al., 2006). ZnCuCoAlLDH exhibit characteristic type IV isotherms (Figure 3), which are typical for mesoporous materials, with high BET surface) and average pore diameter higher than 300 for all samples. The hysteresis loop is of H3 type, which is usually associated with the condensation-evaporation phenomena between aggregates of platelet particles giving rise to slit-shaped pores.

Table 1. Textural and structural characteristics for the samples ZnCuCoAl series obtained by co-precipitation Molar ratio Sample Zn ZCuCoA 10 ZCuCoA 82 ZCuCoA 55 ZCuCoA 28 ZCuCoA 01 1.0 1.0 1.0 1.0 1.0 Cu 1.0 0.8 0.5 0.2 0.0 Co 0.0 0.2 0.5 0.8 1.0 Al 1.0 1.0 1.0 1.0 1.0 LDH LDH LDH LDH LDH () 3.076 3.072 3.072 3.071 3.065 () 22.59 22.70 22.68 22.68 22.86 Crystalline phase a c D(003) (nm) 19.9 17.5 14.5 15.1 16.1 D(110) (nm) 13.8 14.5 20.3 22.5 28.6 SBET (m .g ) 77 96 102 102 93
2 -1

Vt (cm .g ) 0.72 0.83 0.90 0.79 0.72


3 -1

APS () 373 349 353 310 311

The characterization of as-synthesized materials was further accomplished by FTIR spectroscopy to analyze the substitution of the two cations in the ZnAl matrix. The corresponding

FTIR spectra are shown in Figure 4. Beside the three groups of absorption bands at 1630, 1480 and 1357 cm-1 discussed elsewhere (Dragoi et al., 2010a), the spectra show absorption bands below 1000 cm-1 which indicates the formation of crystalline networks containing transitional metals. Generally, these bands are associated with the stretching and deformation modes in M-O, M-O-M and O-M-O specific groups in the brucite-like layers, together with other modes of the carbonate anion (Chang et al., 2005; Rives et al., 2003; Holgado et al., 2001).

Figure 4. FT-IR spectra for as-made ZnCuCoAl LDHs obtained by co-precipitation

Figure 5 presents the DR UV-Vis spectra for the layered double hydroxides of the ZnCuCoAl series. Because Zn2+ and Al3+ have d10 and d0 electronic configurations, they do not show absorption bands in the UV-Vis spectra. Therefore, the spectra from Figure 3 show bands associated with d-d transitions of Cu2+ and Co2+ cations in octahedral coordination in the brucitelike layer. The position of the absorption bands depends on the nature of cations while their intensity depends on their concentrations. As consequence, the absorption band at 528 nm is more intense as the content of substituted Co2+ increases. As cobalt is gradually replaced by copper, the intensity of this band decreases until its disappearance along with an increase in the intensity of a band centered at ~750 nm. For the Co-free sample (ZCuCoA 10) only a band at ~750 nm was

observed, which is attributed to the Cu2+ cations in an octahedral coordination and situated on the edges of the brucite like-layers.

Figure 5. DR UV-Vis spectra for as-made ZnCuCoAl LDHs obtained by co-precipitation

To confirm that the band at ~750 nm is indeed characteristic of Cu2+ cations, selected samples were reduced at 150C under hydrogen flow and their corresponding spectra were recorded (Figure 5). It can be observed that this band completely disappeared and a new band become visible at ~600 nm. Moreover, the intensity of this last band increases with the Cu content. According to literature, the band at ~600 nm is attributed to absorption of the Cuo atoms (Sojka et al., 2008). It is interesting to observe that the adsorption bands at ~240 nm attributed to octahedral Cu2+ are still visible after reduction at 150C, suggesting that only copper cations placed on the edges of the brucite layers and on the crystal defects are reducible under these conditions.

Figure 6. DR UV-Vis spectra for ZnCuCoAl LDHs after reduction at 150C.

The reducibility of Cu2+ and Co2+ in ZnNiCoAl-LDH precursors were investigated by means of TPR. Figure 7 illustrates the TPR curves for two selected samples with different Cu2+/Co2+ molar ratio. Because Zn2+ and Al3+ cations are not reducible in the studied temperature range 25 800 C, the TPR curves reflect the reduction of cobalt and nickel cations to the corresponding zero valent states. For the Cu-rich LDH (sample ZCuCoA 82), a first reduction main peak at ~261C together with a shoulder at ~236 C can be observed, which may be assigned to the reduction of Cu2+ species in two different chemical environments and/or locations in the brucite-like layers (Velu et al., 2001). According with the DR UV Vis results (Figure 5), the low temperature peak can be tentatively attributed to the reduction of copper cations in highly defective positions in the brucitelike layers and may explain the formation of Cu0/LDH nanocomposites by mild reduction treatment. Beside these peaks, a broad and low intense peak can be observed centered at ~ 521 C, which is associated with hardly reducible Co2+ cations (Velu et al., 2000) (S. Velu, K. Suzuki, M. P. Kapoor, S. Tomura, F. Ohashi and T. Osaki, Chem. Mater., 2000, 12, 719.). The Co-rich LDH (sample ZCuCoA 28) shows peaks at similar reduction temperatures as Cu-rich sample (~245 and ~512 C) but of different contributions at the total hydrogen consumption, in agreement with its higher Co/Cu ratio. It is finally worthy of note that comparing with the Co- and Cu-free LDH samples, the reduction temperatures of both Cu2+ and Co2+ are lower in bi-component LDHs (~ 325 C for Co-

free LDH and > 600 C for Cu-free LDH; not shown here), which indicates the positive effect of the presence of one cation on the reducibility of the other.

Figure 7. TPR curves for as-synthesized ZCuCoA82 and ZCuCo28 samples

3.2. Cinnamaldehyde hydrogenation As already introduced, the main objective of this work was to prepare new catalysts of Me/LDH nanocomposite-type based on the mild reduction of non-noble transition metals incorporated in a ZnAl-LDH matrix. Non-noble metals were chosen due to: (i) their interesting properties of chemo- and region-selectivity for the hydrogenation reactions (Hubaut et al., 1989) and (ii) the predisposition to substitute the noble metals for economic concerns, the non-noble metals are cheaper than the noble ones. The catalytic behavior of these Me/LDH nanocomposites was studied in the liquid phase hydrogenation of cinnamaldehyde.The hydrogenation of cinnamaldehyde proceeds via reaction pathways that involve hydrogenation of C=O and/or C=C groups (Scheme 1), the degree of the hydrogenation and the route depending on the activity and selectivity of the catalyst. Cinnamaldehyde (CNA) hydrogenation on Me/LDH catalysts resulted in three products: cinnamyl

alcohol (CNOL) (route 1), hydrocinnamaldehyde (HCNA) (route 2) and hydrocinnamyl alcohol (HCNOL) (routes 3 and 4).

Scheme 1. Reaction pathways for CNA hydrogenation Figure 8 illustrates the catalytic activity of Me/LDH nanocomposites obtained by the mild reduction at 150 oC of as-synthesized ZnCuCoAl LDH precursors. Under such mild conditions, according to TPR data, only Cu2+ are susceptible to be reduced to Cu0 while cobalt remains in cation state due to its low reducibility. It is thus expected that activity mainly depends on the content of Cu in the LDH precursors. Indeed, it can be observed in Figure 8 that after 6 h of reaction the total conversion of CNA drastically increased from ~1 % to 98.53 % for Cu-free and Co-free LDH, respectively. Moreover, the activity gradually increases with the increase in the Cu/Co ratio.

Fig. 8. Hydrogenation of trans-cinnamaldehyde over Me/LDHs nanocomposites derived from ZnCuCoAl LDH precursors . (Reaction conditions: T = 150 C, 0.265 g catalyst; CNA = 1 ml, solvent = propylene carbonate (25 mL), rate stirring = 900 rpm, H2 flow = 1L.h-1).

The high activity of Me/LDH nanocomposites in hydrogenation can be associated without any doubt with the concentration of copper catalytic active species, in particular the types present on the surface of nanocomposites. It is considered that the reduction of copper cations placed in the most favorable positions such as edges and vertices of LDH crystals, crystal defects, etc leads to highly active catalytic active centers. As discussed in our previous study (Dragoi et al., 2010b), copper cations are statistically distributed in the LDH framework and thus, when some of the Cu-OM bonds (M = Co, Mg, Al) are broken by mild reduction (see Figure 1), Cu0 will be highly dispersed in the LDH component of the nanocomposite. Consequently, our data confirm that metallic copper is catalytically active in hydrogenation only when it is in a highly dispersion state (i.e., high electron densities) (Maki-Arvela et al., 2005). As concerns the influence of Co2+, it seems that its main role is to improve the reducibility of Cu2+ cations, since the selectivity to CNOL does not depend at all on the Co content of nanocomposites (maximum of ~ 45 % at a isoconversion of CNA of ~9 % for all compositions).

Fig. 9. Selectivity vs CNA conversion over Me/LDH nanocomposites obtained from ZNiCoA 82 (A) and ZCuCoA 82 (B) precursors. (Reaction conditions: T = 150 C, 0.265 g catalyst; CNA = 1 ml, solvent = propylene carbonate (25 mL), rate stirring = 900 rpm, H2 flow = 1L.h-1). However, when Cu is replaced with Ni in the ZnAl matrix, the role played by cobalt is totally changed. Figure 9 shows a comparison between the selectivity of catalysts derived from ZNiCoA 82 and ZCuCo 82 LDH precursors, which illustrates that the activity and chemoselectivity
of Me/LDH nanocomposites can be finely controlled by an adequate choice of the transition non-noble metal. Thus, as compared with the ZnCuCoAl system, substitution of Co and Ni in the ZnAl matrix results in a lower catalytic activity (9 vs 55 % CNA conversion after 6 h of reaction) but in an enhanced selectivity to unsaturated alcohol (85 vs 45 % at an isoconversion of CNA of 9 %) (Figure 9A and B).

Conclusions

Experimental data described in the present work provide evidence of the possibility of using non-calcined layered double hydroxide substituted in brucite-like layers with cations of non-noble transition metals (Ni2+, Co2+, Cu2+) as precursors for the composites containing nanosized metal particles in LDH matrix. Their layered structure and the substitution of Cu2+ and Co2+ in the brucitelike layers were proved by various methods (XRD, FT-IR, DR UV-Vis, TPR, etc.). Reduction of some transition metal cations located in favorable positions and their encapsulation in the matrix do not lead to a drastic loss in the layered structure when the activation process is carried out at 150C, under hydrogen flow. At the same time, the resulted nanocomposite materials of metal/LDH type exhibited high catalytic activity in the liquid-phase hydrogenation of cinnamaldehyde. On the one

hand, Cu and Co substitution offered a better catalytic activity in terms of CNA conversion. On the

other hand, the catalyst derived from CoNi-containing LDH was found to be more promising for the
selective production of CNOL. Certainly, the catalytic properties could be tailored by modifying the

chemical composition of LDH precursors. Thus, nanocomposites of this kind could attract much
attention for their potential practical applications as catalytic materials, which are easily synthesized, and usefully for carrying out hydrogenation processes under gentle reaction conditions.

Acknowledgement: The authors (A.C., A.U., B.D., E.D.) acknowledge the financial support from CNCSIS-UEFISCSU (Romania) under PNII-Idei Project No. 485/2009. C. Rudolf acknowledges the

financial support of POSDRU CUANTUMDOC Doctoral Studies for European Performances in Research and Inovation ID 79407 project funded by the European Social Found and Romanian Government. A. Sasu acknowledges the financial support of EURODOC Doctoral Scholarships for research performance at European level" ID 59410 project, finance by the European Social Found and Romanian Government.

References

Benito P., Labajos F.M., Rocha J., Rives V., (2006) Influence of microwave radiation on the textural

properties of layered double hydroxides Microporous Mesoporous Mater., 94, 148; Benito P., Herrero M., Labajos F.M., Rives V., Royob C., Latorre N., Monzon A., (2009), Production of carbon nanotubes from methane. Use of Co-Zn-Al catalysts prepared by microwave-assisted synthesis, Chem. Eng. J., 149, 455462; Cavani F., Trifiro F., Vaccari A., (1991), Hydrotalcite-type anionic clays: preparation, properties and applications. Catal. Today 11, 173301; Castelijins A.M.F.C., Hogeweg J.M., van Nispen S.P.J.M, (1998), Process for the preparation of 3phenylpropanal, USA Patent, No. 5811588; Chang Z., Evans D.G., Duan X., Vial C., Ghanbaja J., Prevot V., de Roy M., Forano C., (2005), Synthesis of [ZnAlCO3] layered double hydroxides by a coprecipitation method under steady-state conditions, J. .Solid State Chem., 178, 27662777; Coq B., Figueras F., (1998), Structure-activity relationships in catalysis by metals: some aspects of particle size, bimetallic and support effects, Coord. Chem. Rev., 178, 17531783; Dragoi B., Ungureanu A., Chirieac Al., Hulea V., Dumitriu E., (2010a), Hydrogenation of Unsaturated Carbonyl Compounds on non-Calcined LDHs. I. Synthesis and Characterization of ZnNiCuAl Hydrotalcite-like Materials, Acta Chim. Slov., 57, 677685;

Dragoi B., Ungureanu A., Meloni D., Casula M., Chirieac A., Sasu A., Solinas V., Dumitriu E., (2010b), Cu, Ni based hydrotalcite-like compounds as catalysts for the hydrogenation of cinnamaldehyde in liquid phase. Part 2: Influence of reaction conditions and chemical composition on the catalytic properties, Environm. Engin. Manag. J., 9, 1203-1210;Finta Z., (2000) J. Mol. Cat. A: Chemical, 161, 149-.1 Gallezot P., Richard D., (1998), Selective hydrogenation of ,-unsaturated aldehydes, Catal. Rev. Sci. Eng., 40, 81126; Halawy S.A., Mohamed M.A., Abd El-Hafez S.F., (1994), Dehydrogenation of isopropyl alcohol over Co-Ni/Mg oxide catalysts, J. Mol. Catal. 94, 191-203; Holgado, M.J., Rives, V., San Romn, M.S., (2001), Characterization of Ni-Mg-Al mixed oxides and their catalytic activity in oxidative dehydrogenation of n-butane and propene, Appl. Catal. A: Gen. 214, 219-228; Hubaut R., Bonelle J.P., M. Daage, (1989), Selective hydrogenation of heavy polyunsaturated molecules on copper-chromium catalysts, J. Mol. Catal., 55, 170-183; Mki-Arvela, P., Hjek, J., Salmi, T., Murzin, D.Yu., (2005), Chemoselective hydrogenation of carbonyl compounds over heterogeneous catalysts, Appl. Catal. A: Gen., 292, , 1-49; Matsushita T., Ebitani K., Kaneda K., (1999), Highly efficient oxidation of alcohols and aromatic compounds catalysed by the Ru-Co-Al hydrotalcite in the presence of molecular oxygen, J. Chem. Soc., Chem. Commun., 3, 265-266. Ponec V., (1997), On the role of promoters in hydrogenations on metals: -unsaturated aldehydes and ketones, Appl. Catal. A-Gen., 149, 2748. Porta P., Morpurgo S., (1995), Cu/Zn/Co/Al/Cr-containing hydrotalcite-type anionic clays, Appl. Clay Sci. 10, 31-44 Rives V., (2001), Layered Double Hydroxides: Present and Future, Nova Science Publishers, New York. Rives V., Dubey A., Kannan S., (2001), Synthesis, characterization and catalytic hydroxylation of phenol over CuCoAl ternary hydrotalcites, Phys. Chem. Chem. Phys., 3, 4826-4836. Rives V., Prieto O., Dubey A., Kannan S., (2003), Synergistic effect in the hydroxylation of phenol, over CoNiAl ternary hydrotalcites, J. Catal., 220, 161171. Sojka Z., Bozon-Verduraz F., Che M., in: G. Ertl, H. Knzinger, F. Schth, J. Weitkamp (Eds.), (2008), Handbook of Heterogeneous Catalysis, 8 Volumes, 2nd Edition, Wiley-VCH Verlag GmbH& Co. KGaA, Weinheim, pp. 1047. Tichit D., Ribet S., Coq B., (2001), Characterization of calcined and reduced multi-component CoNi-Mg-Al-Layered Double Hydroxides, Eur. J. Inorg. Chem., 539-546.

Tessonnier, J.-P. Pesant, L., Ehret, G., Ledoux, M.J., Pham-Huu, C., (2005), Pd nanoparticles introduced inside multi-walled carbon nanotubes for selective hydrogenation of

cinnamaldehyde into hydrocinnamaldehyde, Appl. Catal. A: Gen., 288, 203-210; Ungureanu A., Meloni D., Dragoi B., Casula M., Chirieac Al., Solinas V., Dumitriu E., (2010) Cu,Ni-based hydrotalcite like compounds as catalysts for the hydrogenation of cinnamaldehyde in liquid phase. part 1: synthesis and characterization, Envirom. Eng. Manage. J. Vol. 9, 4, 461-468; Vaccari A., (1998), Preparation and catalytic properties of cationic and anionic clays, Catal. Today 41, 53-71. Vaccari A., (1999), Clays and catalysis: a promising future, Appl. Clay Sci. 14, 161198. Velu S., Suzuki K., Kapoor M. P., Tomura S., Ohashi F., Osaki T., (2000), Effect of Sn Incorporation on the Thermal Transformation and Reducibility of M(II)Al-Layered Double Hydroxides [M(II) = Ni or Co], Chem. Mater., 12, 719730. Velu S., Suzuki K., Hashimoto S., Satoh N., Ohashi F., Tomura S., (2001), The effect of cobalt on the structural properties and reducibility of CuCoZnAl layered double hydroxides and their thermally derived mixed oxides, J. Mater. Chem., 11, 2049-260.

Das könnte Ihnen auch gefallen