Kazem Karami*a,
Sedigheh Abedanzadeha and
Pablo Hervésb
aDepartment of Chemistry, Isfahan University of Technology, Isfahan 84156/83111, Iran. E-mail: karami@cc.iut.ac.ir
bDepartamento de Quimica Fisica, Universidade de Vigo, 36310 Vigo, Spain. E-mail: jherves@uvigo.es
First published on 26th September 2016
The present work describes the preparation of organically modified TiO2-supported Pd catalyst originated from the new benzophenone imine-derived CN-palladacycle. The heterogeneous organic–inorganic hybrid catalyst system has been characterized by FT-IR, XRD, SEM, EDX, TEM and XPS techniques and exhibited good catalytic activity in the Sonogashira cross-coupling reactions of phenylacetylene with aryl halides under copper-, amine- and phosphine-free conditions, in conjunction with the ultra low catalyst Pd-loading. Significantly, the heterogeneous Pd catalyst allowed the reaction of phenylacetylene with aryl iodides to improve in excellent yields under very mild conditions using green solvent. Finally, the reusability of the supported Pd-complex was investigated by multiple reuses of the supported catalyst in subsequent Sonogashira cross-couplings.
The Sonogashira coupling reaction of terminal alkynes with aryl halides is one of the most important methods which has been widely applied for the synthesis of natural products, pharmaceutical compounds and polymeric materials.18–23 This reaction is generally catalyzed by homogeneous Pd complexes in the presence of toxic phosphine ligands, using copper salts as co-catalysts and amine as a solvent or base, under inert condition which has so many drawbacks.24–29 Consequently, there have been many efforts to introduce highly active and easily reusable catalysts through convenient Cu-, amine and phosphine-free methods under environmentally friendly conditions.
Supported palladium catalysts with high dispersion and narrow size distribution of palladium particles would address the limited practical application of homogeneous palladium catalysts and exhibit high activity, recyclability, low palladium loading and promising performances in the synthetic and industrial chemistry. In this regard, a wide variety of solid materials as heterogeneous Pd supports including active carbon,30 mesoporous silica,31 metal oxides,32 zeolites,33 magnetic materials,34 hydrotalcites,35 hydroxyapatite,36 and organic–inorganic hybrid37 have been investigated. However, many heterogeneous catalytic systems suffer from high palladium loading and/or low catalytic activities due to the palladium agglomeration and catalyst deactivation.38
In this study, we report the synthesis and characterizations of a nitrogen-containing orthopalladated complex driving from bis[4-(dimethylamino)phenyl]methaniminium chloride (Auramine-O). In continuation of our interest in developing new supported catalysts,39–42 herein we use a strategy to convert a homogeneous catalyst into a heterogeneous one through the stabilization of active Pd sites onto the large surface area of inorganic support by an organic spacer to create organic–inorganic hybrid catalyst. A simple route is described to prepare a supported CN-palladacycle catalyst based on functionalized TiO2 as a very suitable inorganic support. The resulting ultra Pd-loaded material was investigated as a catalyst for copper- and amine-free Sonogashira coupling reactions of phenylacetylene with aryl halides which were carried out in the absence of phosphine ligands under air. In the context of green chemistry, heterogeneous catalysis of Sonogashira coupling reactions of phenylacetylene with aryl iodides carried out in ethanol as a green solvent.
New dinuclear CN-palladacycle was synthesized through the reaction of palladium(II) acetate with bis[4-(dimethylamino)phenyl]methaniminium chloride (Auramine-O) in 1:
1 molar ratio (Scheme 1). The green orthopalladated chloro-bridged complex could be produced directly from the refluxing reaction of Pd(OAc)2 with Auramine-O ligand in toluene for one 24 hours (method A). However, during the second method (B), the reaction of Pd(OAc)2 with Auramine-O ligand was carried out in THF, at room temperature for higher reaction time of 72 hours to afford the pure product in higher yield. The dinuclear CN-palladacycle was characterized by IR and NMR spectroscopies. The elemental analysis result of the prepared compound was in good agreement with the calculated values.
As illustrated in Fig. 1, the FT-IR spectrum of Auramine-O ligand shows the peak at 1685 cm−1 due to the CN stretching vibration.44 This band appears at the lower energies (1606 cm−1) in the spectrum of synthesized dinuclear complex (Fig. 1b) indicating the coordination of the ligand to the metal. As clearly shown in Fig. 1a, the iminium salt displays strong, broad N–H stretching absorptions in the 2250 to 3000 cm−1 region, where overlap with C–H absorption occurs. The lower frequency region of FT-IR spectrum shows the characteristic bands at 674 and 416 cm−1 related to the ν(Pd–C) and ν(Pd–N) vibrations in the dinuclear complex, respectively (Fig. 1b).45,46 These vibrational bands confirm the successful coordination of the CN-orthopalladation of Auramine-O ligand to the palladium(II) center by covalent interaction.
![]() | ||
Fig. 1 FT-IR spectra of: (a) Auramine-O ligand, (b) dinuclear palladium complex (ADP) (c) TiO2, (d) fresh TiO2-supported Pd catalyst. |
In the 1H NMR spectrum, the signal related to the methyl protons of cyclometallated Auramine-O appears at 3.00 ppm. All of the 1H NMR signals for the aromatic protons of the complex are shifted with respect to the parent iminium salt confirming the coordination of Auramine-O.
In order to design heterogeneous catalysts, solid supports containing nitrogen as donor atoms have been achieved special interest. Since N-donor ligands are air stable, less expensive and nontoxic, they can be used to replace phosphine groups and coordinate to the various transition metal complexes.
Due to the fact that inorganic supports like TiO2 show advantages such as mechanical stability and resistance against solvent, chemical reagents and high temperature,49 a general rout has been designed to support a CN-palladacyclic complex onto the organically modified-TiO2 surface to prepare an efficient heterogeneous pre-catalyst. The synthetic procedure is illustrated in Scheme 2.
![]() | ||
Scheme 2 Synthetic method and proposed route for the synthesis of new TiO2-supported catalyst. (i) Dry toluene, 12 h, reflux (ii) acetone, 24 h. |
TiO2 usually shows pH-dependent surface charges in the aqueous solution because of the existence of Ti–OH on the surface. The absorption property of TiO2 greatly changes with environment of different pH value. Considering that the catalytic properties of supported catalysts depend on their preparation conditions, we tried to prepare the ligand functionalized TiO2 using the pH-dependent procedure.50,51 To obtain the TiO2-supported Pd catalyst, a mixture of chloro-bridged orthopalladated complex (ADP) and a functionalized TiO2 were refluxed in acetone for 24 hours.
The stabilization of CN-palladacycle onto the functionalized TiO2 is due to donation of an electron lone pairs on the nitrogen groups of 3-aminopropyltrimethoxysilane (APTES) into the unoccupied orbital of the palladium atoms in continuous of the bridge cleavage reaction of the dimeric compound ADP (Scheme 2).
Fig. 1 shows the FT-IR spectra of Auramine-O ligand, Auramine-O dimeric palladacycle (ADP), TiO2 and fresh TiO2-supported Pd catalyst. Corresponding to the FT-IR spectrum of TiO2-supported catalyst illustrated in Fig. 1d, the broad band centered at 500–600 cm−1 is likely due to the vibration of the Ti–O bonds in the TiO2 lattice.52 The peak at 1627 cm−1 and the broad peak appearing at 3451 cm−1 are assigned to vibrations of hydroxyl groups.45,52,53 In modified sample these peaks overlap with broad bands of the stretching and deformation modes of NH2 groups.45,54 The stretching vibrations of aliphatic groups are appeared at 2951 cm−1. The bands at 1124 and 1383 cm−1 are attributed to the characteristic absorptions of the Si–CH2 and Si–O groups, respectively.55 Also, the presence of several bands with medium intensity in the 2839–3062 cm−1 regions is allocated to C–H stretching of groups (symmetric and asymmetric stretching).
The elemental analysis further proved the existence of C, H, and N atoms in the prepared TiO2-supported Pd catalyst.
Powder X-ray diffraction (XRD) analysis for fresh catalyst exhibits the sharp reflections belonging to the TiO2 nanoparticles in the anatase phase (Fig. 2). Although no Pd species could be detected by XRD pattern due to the low amount of Pd content or its high dispersion on the solid surface, the presence of Pd species was clearly evidenced by X-ray photoelectron spectroscopy (XPS) verifying the orthopalladated complex ADP coated on the TiO2.
Fig. 3 shows high resolution XPS spectra of TiO2-supported Pd catalyst. The Pd 3d spectrum in Fig. 3a related to the fresh catalyst, shows the characteristic Pd 3d5/2 and 3d3/2 doublet peaks, with a peak separation of 5.26 eV. The fitted 3d5/2 peaks are located near the expected positions for the Pd0 and Pd2+ species at around 335.3 (ref. 56–58) and 337.6 eV,56,57 respectively. The fitting indicates that Pd in the fresh catalyst is a mixture of Pd2+ and Pd0 states. After the first run, Pd0 is observed in higher proportions, which indicates that after the catalytic cycle Pd2+ cations were transformed into Pd0 (Fig. 3b). According the Pd 3d spectrum of Pd catalyst after the forth catalytic cycle (Fig. 3c), presence of the characteristic Pd 3d5/2 and 3d3/2 doublet peaks related to the Pd0 species confirming the conversion of the whole Pd2+ ions to Pd0. In fact, the percentage of Pd2+ decreases from 53% for fresh catalyst to 36% from first run and approximately 0% for the fourth run as we can see in Fig. 3. According to the XPS data of Pd catalyst before and after the recycle, the Si 2p spectrum shows only one component, with the Si 2p peak position around 103.3 eV, assigned to Si–O bonds.58 The Ti 2p spectrum shows the characteristic Ti 2p3/2 and 2p1/2 doublet peaks, with a peak separation of 5.54 eV. The characteristic shake-up peaks for TiO2 are also observed. The fitted 2p3/2 peak is located near the expected positions for the TiO2 species at around 458.7 eV.58 The HR-XPS N 1s spectra of the fresh Pd catalyst shows two components, one at around 400.0 eV attributed to C–N bonds58 and another minor component at around 401.9 eV attributed to N+–H bonds.59 The N 1s fitted spectrum after the first recycle shows one component at around 400.0 eV attributed to C–N bonds58 and after the forth recycle, the nitrogen presented a very low signal and the peak was not possible to fit. The Cl 2p spectrum of the fresh Pd catalyst shows the characteristic Cl 2p3/2 and 2p1/2 doublet peaks, with a peak separation of 1.6 eV. After the recycle, no chlorine signal is detected which indicate that after the catalytic cycles Pd2+ species were transformed into Pd0 and the ligand was removed from palladacycle.
![]() | ||
Fig. 3 XPS spectra of high resolution Pd 3d of (a) fresh TiO2-supported Pd catalyst, (b) after the first run and (c) after the forth run. |
As illustrated in the similar investigations60 the formation of TiO2-supported Pd(0) nanoparticles are possibly due to the donation of an electron lone pairs on the nitrogen groups into an unoccupied orbital of the palladium nanoparticles. Moreover, the presence of ethanol as solvent, possibly leads to the reduction of Pd(II) to Pd(0).
XPS analysis showed that the palladium content of fresh catalyst was about 1.5%. The palladium content of catalyst after the reaction cycle was considerable, according to the XPS analysis. Further information about homogeneous or heterogeneous nature of the catalyst was obtained by hot filtration test which were in accordance with the results obtained from XPS analysis and confirmed that using this catalyst reaction proceeds mostly under heterogeneous conditions.
To obtain the morphology and particle size of the catalyst, FE-SEM images of the catalyst were carried out (Fig. 4). Furthermore, energy-dispersive X-ray spectroscopy (EDX) obtained from SEM shows the presence of Pd atoms as well as Ti, Si, N, and O atoms in the structure of fresh TiO2-supported Pd catalyst (Fig. 5).
According to the TEM images, the Pd nanoparticles are formed and well dispersed through the supported inorganic surface. It was verified that synthesis of catalyst through the pH adsorption method played an important role in the distribution of nanoparticles. The size distribution of fresh TiO2-supported Pd catalyst is shown in Fig. 6. The histogram could be fitted to a Gaussian function, yielding a mean size of 2.24 nm based on the sizes of 90 particles (Fig. 6a). According to the TEM images after the first run, the average particle size of Pd0 in supported Pd catalyst was 2.61 nm. The average particle size of Pd0 after the forth run was 3.31 nm.
![]() | ||
Fig. 6 TEM images of (a) fresh TiO2-supported Pd catalyst, (b) after the first run and (c) after the forth run. The histograms illustrate the size distribution of Pd catalyst. |
The effectiveness of supported Pd catalyst for the Sonogashira cross-coupling reaction of iodobenzene with phenylacetylene was initially investigated. First we investigated the effect of solvent in the Sonogashira reaction of iodobenzene with phenylacetylene (Table 1, entry 1–7). The yield was significantly decreased when H2O was used as the solvent (Table 1, entry 5); this may be due to the insolubility of the substrate and catalyst in water. However, application of 1:
1 aqueous EtOH improves the yield to some extent (Table 1, entry 6). On the other hand inferior results were obtained with solvents like THF, CH3CN, toluene and DMF (Table 1, entries 1–4). As shown in Table 1 better performance of the catalytic system was observed when EtOH was used and based on the above we have considered EtOH for further optimization. A very low Pd-loading of 0.005 mol% was found to be the optimum amount of catalyst. To investigate the effect of the base which plays a crucial role in the overall performance of a catalytic system in a Sonogashira reaction, several inorganic and organic bases were tested (Table 1, entry 7–10). K2CO3 was found to be reactive in our catalytic system and we have considered it for further optimization. The preliminary investigation for reaction optimization was performed with 0.005 mol% of Pd catalyst and two equivalent of K2CO3 in EtOH (4 ml) at 85 °C under aerobic condition. The results obtained are highlighted in Table 1. The turnover frequency of the catalyst was defined to be 5880, which was regarded as a factor to show the high efficiency of the catalyst.
Entry | [Pd] (mol%) | Solvent | Base | Time | Temperature (°C) | Conversiona (%) |
---|---|---|---|---|---|---|
a Reactions were monitored using GC. | ||||||
1 | 0.005 | DMF | K2CO3 | 6 h | 85 | 60 |
2 | 0.005 | CH3CN | K2CO3 | 6 h | 85 | 41 |
3 | 0.005 | Toluene | K2CO3 | 4 h | 85 | Trace |
4 | 0.005 | THF | K2CO3 | 4 h | 85 | Trace |
5 | 0.005 | H2O | K2CO3 | 4 h | 85 | Trace |
6 | 0.005 | Ethanol![]() ![]() ![]() ![]() |
K2CO3 | 4 h | 85 | 70 |
7 | 0.005 | Ethanol | K2CO3 | 30 min | 85 | 99 |
8 | 0.005 | Ethanol | Na2CO3 | 6 h | 85 | 43 |
9 | 0.005 | Ethanol | K3PO4·3H2O | 6 h | 85 | 90 |
10 | 0.005 | Ethanol | NaOAc | 4 h | 85 | 10 |
11 | 0.005 | Ethanol | K2CO3 | 6 h | R. T. | Trace |
12 | 0.02 | Ethanol | K2CO3 | 15 min | 85 | 99 |
To demonstrate the versatility of the catalytic system, we investigated the reaction using a variety of aryl iodide with phenylacetylene under the optimized conditions. It was observed that aryl iodides underwent cross coupling within a short reaction time at very low catalyst Pd-loading of 0.005 mol%, with excellent yields (Table 2). They may produce very high TOFs in such catalytic reactions.
Entry | R | Time | Conversion (%) | TONb | TOFc (h−1) |
---|---|---|---|---|---|
a Reaction conditions: aryl halides (1.0 mmol), phenylacetylene (1.2 mmol), K2CO3 (2.0 mmol), EtOH (4 ml), catalyst (0.005 mol% [Pd]), 85 °C. Reactions were monitored using GC.b TON = mmol of products/mmol of catalyst.c TOF = TON/time. | |||||
1 | H | 30 min | 98 | 5880 | 5880 |
2 | p-NO2 | 20 min | 99 | 5940 | 11![]() |
3 | m-NO2 | 1 h | 90 | 5400 | 2700 |
4 | p-MeO | 2 h | 99 | 5940 | 5940 |
5 | m-MeO | 3 h | 99 | 5940 | 1980 |
To study the performance of TiO2-supported Pd catalyst for the Sonogashira cross-coupling reaction of aryl bromide with phenylacetylene, optimization of the reaction conditions with various solvents and catalyst dosages were applied for bromobenzaldehyde in different range of time reactions and temperatures (Table 3). Among the different reaction conditions, maximum conversion was observed when we have considered the reaction in DMF (4 ml), in the presence of 0.02 mol% of Pd catalyst at 110 °C, under aerobic condition for 8 hours.
The generality of the current system has been investigated with several electronically diverse aryl bromide and phenylacetylene (Table 4, entries 1–8). Different functionality at the phenyl ring of the aryl halide moiety affects the reaction conditions. When the reactions of activated aryl bromides containing electron deficient phenyl rings with phenylacetylene were carried out in DMF at 110 °C for 8 h, excellent yields of desired cross-coupling products were observed. However, the electron donating groups at the para position of aryl bromides couple smoothly with phenylacetylene to give good to excellent yield of cross coupling products. Considering the steric effect, aryl bromides with ortho substituents, provided the corresponding products in high yields (Table 4, entry 4).
Entry | R | X | Conversionb (%) | TON | TOFc (h−1) |
---|---|---|---|---|---|
a Reaction conditions: aryl halides (1.0 mmol), phenylacetylene (1.2 mmol), K2CO3 (2.0 mmol), DMF (4 ml), catalyst (0.02 mol% [Pd]), 110 °C. Reactions were monitored using GC and TLC.b TON = mmol of products/mmol of catalyst.c TOF = TON/time. | |||||
1 | H | Br | 50 | 750 | 93.75 |
2 | p-CHO | Br | 95 | 1425 | 285 |
3 | p-MeCO | Br | 94 | 1410 | 235 |
4 | o-NO2 | Br | 97 | 1455 | 181.875 |
5 | m-Me | Br | 10 | 150 | 18.75 |
6 | 1-Bromonaphthalene | 30 | 450 | 56.25 | |
7 | 2-Bromopyridine | 50 | 750 | 93.75 | |
8 | 2-Bromothiophene | 57 | 855 | 106.875 | |
9 | H | Cl | 50 | 750 | 93.75 |
10 | p-NO2 | Cl | 62 | 930 | 116.25 |
11 | p-NH2 | Cl | 28 | 420 | 52.5 |
12 | o-OH | Cl | 34 | 510 | 63.75 |
Finally Sonogashira reactions of phenylacetylene and aryl chlorides were also investigated in our present system, under the optimized condition used for the case of aryl bromides. To check the effectiveness of our protocol, phenylacetylene was reacted with electron rich 4-aminobenzene and electron deficient nitrochlorobenzene. Although aryl chlorides are not so reactive and are employed less in palladium-catalyzed coupling reactions,63–65 modest to good yield of desired cross coupling products was observed in all cases (Table 4, entries 9–12).
A possible catalytic reaction cycle for the Sonogashira reaction has been shown in Scheme 3, over the active Pd-sites present in the TiO2-supported heterogeneous catalyst. At first, oxidative addition of aryl halide to the Pd catalyst takes place, followed by transmetallation of alkyne and lastly, reductive elimination of the final product.
![]() | ||
Scheme 3 Possible catalytic cycle of the heterogeneously catalyzed Cu-free Sonogashira reaction over TiO2-supported Pd catalyst. |
Aryl halides containing electron-withdrawing groups improve the oxidative addition of C–X to the Pd catalyst, as an activated halo derivatives.25 However, electron-rich aryl halides require harsher experimental conditions to give acceptable results. Although, the mechanism of the Sonogashira reactions has not yet been established clearly, we assume that the lower tendency of electron-rich aryl halides towards metallo-oxidative insertions could be responsible for this diminished reactivity.66,67
Although, there is no second metal in the copper-free Sonogashira reaction, the second step is conventionally called transmetallation. Several studies have attempted to understand how the alkyne is transferred to the Pd catalyst through the transmetallation step. Amines generally used in the Sonogashira reaction, appreciably deprotonated the alkyne in solution. But despite the fact that the Sonogashira reaction is occurred in the amine-free condition in this study, we cannot have the deprotonation of the alkyne in solution. Therefore, deprotonation is hypothesized to occur on the Pd center. Initially, the alkyne coordinates to the Pd center in the η2-fashion, pulling electron density away from the acetylene. This makes the terminal proton more acidic, enabling deprotonation by weaker bases (Scheme 3).68–70 Importantly, the electronic properties of the terminal alkyne have considerable effects the on the deprotonation mechanism: electron withdrawing groups (EWG) decrease and electron donating groups (EDG) increase the energy required for deprotonation.68 The base initially forms the more reactive species of the type Pd0L2 in the oxidative addition step lead to accelerate the overall reaction.70 Using of stronger bases, an excess amount of phenylacetylene and more acidic phenylacetylenes considerably increase the rate by affecting the deprotonation step.71
Furthermore, the reaction kinetics of Sonogashira reaction of phenylacetylene with iodobenzene with 0.02 mol% catalyst was investigated and the result is shown in Fig. 7. A control experiment indicated that the coupling reaction did not occur smoothly in the absence of a catalyst. The yield of the product increased quickly with reaction time until it reached 96% at the 30 minutes.
Entry | Catalysta | Temp. (°C)/time (h) | Medium | Yield (%) | Ref. |
---|---|---|---|---|---|
a Data in parentheses indicate mol% of Pd used. | |||||
1 | MMT@Pd/Cu (1) | 65/24 | DME/H2O | 97 | 72 |
2 | Pd@Fe3O4 (0.5) | 110/24 | DMF | 84 | 73 |
3 | PNP–SSS (1.2) | 100/3 | H2O | 95 | 74 |
4 | FDU–NHC/Pd(II) (1) | 100/3 | DMA | 94 | 75 |
5 | CS/MMT/Pd (0.3) | 110/5 | DMSO | 94 | 76 |
6 | Si–P4VPy–Pd (0.5) | 120/0.25 | NMP | 90 | 77 |
7 | Pd/APTES/TiO2 (0.005) | 85/0.5 | EtOH | 99 | This study |
![]() | ||
Fig. 8 A recyclability test for catalyst Sonogashira cross-coupling reaction of iodobenzene with phenylacetylene (reaction condition are the same as given in Table 2). |
Yield (62–76%); anal. calc. for C34H24N6 Cl2Pd2: C, 51.03; H, 3.02; N, 10.50; found: C, 51.87; H, 2.98; N, 10.28; IR (KBr, cm−1): ν(Pd–N) = 416, ν(Pd–C) = 674, ν(CN) = 1606; 1H NMR (DMSO-d6, ppm): δ = 3.00 (d, 24H, aliphatic), 6.38 (d, 2H, aromatic), 6.81 (d, 4H, aromatic), 7.01 (d, 2H, aromatic), 7.41 (d, 4H, aromatic), 7.47 (s, 1H, aromatic), 8.94 (s, 1H, aromatic).
This journal is © The Royal Society of Chemistry 2016 |