Elisabetta Grazia
Tomarchio
ab,
Chiara
Zagni
*a,
Vincenzo
Paratore
c,
Guglielmo Guido
Condorelli
c,
Sabrina Carola
Carroccio
d and
Antonio
Rescifina
a
aDepartment of Drug and Health Sciences, University of Catania, V.le A. Doria 6, 95125, Catania, Italy. E-mail: chiara.zagni@unict.it
bDepartment of Biomedical and Biotechnological Sciences, University of Catania, Via Santa Sofia 97, 95123 Catania, Italy
cDepartment of Chemical Science, Università degli Studi di Catania, Viale Andrea Doria 6, 95125, Catania, Italy
dInstitute for Polymers, Composites, and Biomaterials CNR-IPCB, Via Paolo Gaifami 18, 95126, Catania, Italy
First published on 26th September 2024
Biobased catalysts play a crucial role in sustainable chemistry, using natural resources to support eco-friendly processes. While palladium catalysts are essential for various industrial applications, they often pose environmental challenges due to their non-reusability and tendency to degrade. To address these issues, we developed an innovative phenylalanine-based catalyst containing palladium (C-PhebPd) designed for the Suzuki–Miyaura reaction. The natural amino acids, used as monomers, chelate palladium, preventing leaching, unlike other heterogeneous catalysts that use palladium nanoparticles, which can be released over time, leading to catalyst degradation. Such catalyst exhibits outstanding performance in aqueous media at moderate temperatures, facilitating cross-coupling reactions between various aryl halides and arylboronic acids with high yields of up to 99%. The affordable synthetic procedure and C-PhebPd's stability make it potentially scalable for industrial applications. The robustness of this catalyst was also proved by recyclability tests up to seven cycles. Further investigation into its capabilities could unlock additional insights for various catalytic transformations.
Efficient separation and recycling of catalyst systems remain ongoing challenges crucial for economic and ecological reasons. From a pharmaceutical perspective, ease of product isolation and minimizing residual impurities are essential factors. Moreover, homogeneous catalysts often rely on harmful organic solvents for optimal reactivity. This leads to diminished performance in the “Do Not Significantly Harm” (DNSH) principle and recycling difficulties due to palladium leaching.8
Given the challenges associated with homogeneous catalytic systems, such as palladium leaching and the reliance on toxic organic solvents, there is significant interest in developing efficient heterogeneous catalysts.9,10 These catalysts are actively being researched for their potential to offer robust reusability and enhance environmental sustainability.11 The goal is to create cleaner and more efficient catalytic systems that eliminate the need for harmful solvents and improve overall environmental impact. For this purpose, several heterogeneous catalysts have been developed, which show prominent advantages such as ease of handling, recyclability, and broad substrate scope.12 Most of the previous heterogeneous catalysis involving palladium metal is transitioning into nanoparticle exploitation.13–17 However, significant drawbacks constrain a wide application due to their high costs, elevated temperatures and pressure required for the process, generation of numerous side products, and management of the catalyst in leaching and recovery. In recent decades, Suzuki–Miyaura coupling reactions carried out on supported Pd nanoparticles provided outstanding ability in converting the reactants. Various porous materials like silica-supported salts,18–20 zeolites,21 carbon-based materials,22 and porous organic polymers (POPs)23 have been utilized as efficient supports.24,25 However, mechanistic studies have shown that using supported Pd nanoparticles as catalysts causes the release of Pd atoms from the catalyst surface during the oxidative addition of Pd(0) species with aryl halides. This leads to the formation of soluble Pd(II) complexes that promote the reaction through a (quasi)homogeneous mechanism. Therefore, preventing catalytic deactivation and product contamination from Pd leaching is difficult.26 Unlike supported Pd nanoparticles, anchoring Pd complexes onto solid supports via well-defined coordination sites is advantageous for producing highly dispersed and stabilized Pd species. This approach helps preserve catalytic stability throughout the reaction and mitigates leaching phenomena. Therefore, in light of these considerations and given a transition towards greener chemistry practices, we synthesized a novel macroporous biopolymeric material derived from phenylalanine complexed with palladium ions as a heterogeneous catalyst. The new material was synthesized by radical polymerization at subzero temperatures in the presence of water.27,28 This technique allows the production of a sponge-like material with high porosity, stability, and good absorption capacity.29 Within the pores of these cryogels, organic molecules become concentrated, facilitating efficient reactions. This innovative approach, which utilizes a Palladium complex, offers promising applications in various catalysis fields and provides several benefits, including affordable natural feedstocks, high product yields, short reaction times, excellent atom economy, and recyclability.
1H NMR (500 MHz, DMSO-d6) δ 7.35 (d, J = 7.8 Hz, 2H), 7.20 (ddd, J = 22.4, 14.9, 7.9 Hz, 7H), 6.68 (dd, J = 17.6, 10.9 Hz, 1H), 5.77 (d, J = 17.7 Hz, 1H), 5.20 (d, J = 11.0 Hz, 1H), 3.73 (d, J = 13.7 Hz, 1H), 3.55 (d, J = 13.7 Hz, 1H), 3.22 (d, J = 6.6 Hz, 1H), 2.92 (dd, J = 13.6, 5.9 Hz, 1H), 2.80 (dd, J = 13.6, 7.4 Hz, 1H).
13C-NMR (126 MHz, DMSO-d6): δ = 174.67, 139.63, 138.73, 136.95, 129.81, 128.87, 128.51, 126.47, 114,37, 62.59, 50.86, 40.08.
1H NMR (500 MHz, DMSO-d6) δ 7.55–7.41 (m, 4H), 7.28–7.22 (m, 5H), 6.75 (dd, J = 17.6, 10.9 Hz, 1H), 6.12 (m, J = 4.7 Hz, 1H), 5.88 (d, J = 17.7 Hz, 1H), 5.31 (d, J = 11.1 Hz, 1H), 3.79 (dd, J = 13.4, 5.0 Hz, 1H), 3.47 (dd, J = 13.4, 3.5 Hz, 1H), 3.29 (d, J = 6.0 Hz, 1H), 3.14 (dd, J = 14.2, 6.0 Hz, 1H), 3.02 (dd, J = 14.2, 6.4 Hz, 1H). 13C-NMR (125 MHz, DMSO-d6): δ = 180.01, 137.33, 137.02, 136.14, 133.64, 133.64, 130.73, 129.40, 128.33, 126.59, 126.25, 114.87, 63.56, 53.54, 37.71.
Subsequently, 1.5% v/v of 10% APS and TEMED ready solutions were added to the mixture. After stirring for 1 minute, the solution was transferred into a 10 mL capped syringe and placed in a cryostatic bath at −15 °C for 24 hours. The resulting cryogel was then thawed, rinsed with water and ethanol, and dried, yielding an 84% polymerization efficiency.
The catalyst's thermal stability was assessed using a PerkinElmer thermogravimetric apparatus under a nitrogen atmosphere (flow rate 60 mL min−1), employing a heating rate of 10 °C min−1 from 90 °C to 700 °C. The TGA sensitivity was 0.1 μg, with a weighting precision of ± 0.01%.
The macroporous morphology of the synthesized material was confirmed by scanning electron microscopy (SEM) with a Phenomenex microscope. Before testing, samples were coated with gold to enhance conductivity. Images were captured to assess the cryogel morphology, with data acquisition and processing conducted using Phenom Porometric 1.1.2.0 software by Phenom-World BV, Eindhoven, The Netherlands. Furthermore, the elemental composition of the catalyst was measured by energy-dispersive X-ray spectroscopy (EDX).
Swelling tests were carried out on dried bare cryogels, measuring mass before and after water uptake. Total uptake was determined by observing cumulative mass increase over fixed intervals.33 Equilibrium swelling was assessed by weighing wet samples immersed for 30 minutes. Adsorption kinetics were analyzed by exposing samples to excess water for specific durations and rapidly removing unabsorbed water. The adsorbed water was quantified by weighing samples over time and normalizing the data. Three parallel samples were tested, and the average result was computed. The standard deviation was found to be less than 5%.
X-ray photoelectron spectroscopy (XPS) was carried out with a PHI 5000 Versa Probe Instrument (Chanhassen, MN, USA) using a monochromatic Al Kα X-ray source excited with a micro-focused electron beam. All the analyses were performed with a photoelectron take-off angle of 45° (relative to the sample Surface). The XPS binding energy (B.E.) scale was calibrated on the C1s peak of adventitious carbon at 285.0 eV.
A sample of 5 mg of C-PhebPd was poured into a vial containing 5 ml of water to determine the possible Pd leaching. After 24 h, 100 μL of the sample was taken for analysis to allow it to reach the equilibrium.36
L-Phenylalanine (L-Phe) contains a nonpolar phenyl group and is a prevalent aromatic amino acid in living organisms.39 Owing to their oxygen and nitrogen atoms, two L-Phe molecules can interact with palladium, forming a stable complex.40 In the pursuit of synthesizing biobased materials, L-Phe underwent functionalization through a reaction with 4-vinylbenzyl chloride in the presence of a base at room temperature (Fig. 1). Initially, the reaction was carried out in methanol, which produced the desired product within two days but also led to the formation of side products. To overcome this issue, the solvent was switched to water. While this change extended the reaction time to three days, it enabled the isolation of pure compound (Pheb) through straightforward filtration. Additionally, this approach aligns with the DNSH principle, providing further environmental advantages. The obtained monomer was mixed with palladium acetate in a 1:1 water/acetone mixture to form the final complex PhebPd 4.
PhebPd 4 underwent radical polymerization in the presence of MBAA at subzero temperature, achieving a grey-coloured macroporous biopolymer-based PhebPd cryogel (Fig. 2).28 The hydrophilic material obtained possesses high swelling and absorption capacity, providing an ideal space to confine organic molecules that favour interactions after solution uptake. Consequently, it is reasonable to suppose this could facilitate intimate contact between the metal and reactants, enhancing the catalytic efficiency.41
FTIR spectra of monomer and polymer were carried out to confirm further the formation of the palladium complex and the polymerization that occurred (Fig. 3(a)). The spectrum of Pheb 3 (black curve) shows an intense peak at 1572 cm−1 related to the asymmetric stretch of the carboxylate group of the amino acid derivative in zwitterionic form, with the amine exhibiting a weaker signal around 1497 cm−1. In the PhebPd spectrum (red curve), a shift of the two signals to 1685 cm−1 and 1560 cm−1 can be observed, corresponding to the stretching of CO and NH deformation, respectively, indicating coordination with Pd. This is further corroborated by the shift in the asymmetric stretching of NH at 3185 cm−1. Additionally, the difference between the symmetric and asymmetric stretching of the carboxylate group (C–O–Pd and CO at 1658 cm−1 and 1352 cm−1, respectively) can confirm the formation of a monodentate bond with Pd.42 In the blue curve, representing the FT-IR analysis of C-PhebPd, a strong signal at 1700 cm−1 is observed, related to the stretching of CO from the carboxylic groups present in the polymer. Peaks corresponding to C–O–Pd, NH–Pd, and NH stretching, similar to the spectrum of PhebPd, are also observed. These peaks in C-PhebPd exhibit no significant shifts but have lower intensities due to the mixture in the polymer matrix. Across all curves, a signal around 3400 cm−1, related to OH stretching from water absorption, is present.
Fig. 3 (a) FTIR spectra of Pheb 3, C-Pheb, and PhebPd; (b) thermogravimetric analysis (TGA) of C-Pheb and C-PhebPd. |
Thermogravimetric analyses (TGA) were conducted to examine the synthesized cryogels’ thermal stability and the residue's Pd content (Fig. 3(b)). The TG curve of C-PhebPd exhibits two distinct weight loss stages. The first stage in the temperature range of 220–240 °C shows a sharp weight loss of 20%, corresponding to the Phe degradation.43 As the temperature rises, a series of reactions occur, causing dehydration, chain cleavage, and depolymerization, with the most significant loss of weight observed. Notably, the residual part obtained at 600 °C is significantly higher than C-Pheb (10%) due to the presence of the metal residue.44
The surface morphology of the new material was investigated using SEM and XPS. Scanning electron microscopy image illustrating the C-PhebPd catalyst revealed a morphology characterized by a macroporous structure with pore diameter ranging from 10 to 60 μm (Fig. 4 and Fig. S10, ESI†). Additionally, the elemental compositions of C-PhebPd material were ascertained using energy-dispersive X-ray (EDX) analysis. As expected, the peak around 3 keV (Fig. S11, ESI†) confirms the presence of Pd, which is also visible to the necked eye, as seen in the grey-coloured cryogel in Fig. 4(b). From the EDX analysis, the weight concentration of Pd content in the polymer was determined to be 9.6%, which agrees with the TGA data.
Fig. 4 (a) SEM image of C-PhebPd; (b) representative image of C-PhebPd; (c) table with element composition of C-PhebPd. |
The pore properties obtained through SEM image analysis are reported in Table S1 (ESI†).
The XPS analysis was performed to support FT-IR observations regarding the formation of the complex and the complementary NMR investigation. The study shows the formation of the palladium complex with PheB (Fig. 5). As depicted in Fig. 5(a), the C1s signal shows three components: a main component at 285 eV due to the typical carbon atoms in hydrocarbon configuration (C–C/C–H), a second component at 288,1 eV due to the presence of carboxylate groups and another component assigned to carbon atoms bonded to one oxygen atom or nitrogen (C–O/C–N) at 286.6 eV. The O1s region shows a main peak at 531.0 eV assigned to the oxygen of COO− groups (Fig. 5(b)). Additionally, the N1s signal (Fig. 5(c)) exhibits a narrow peak at 401.4 eV, associated with –N+–H groups, confirming the presence of derived amino acid molecules. In Fig. 5(d)–(g), the signals related to C1s, O1s, N1s, and Pd3d are reported, confirming the formation of the metal–organic palladium complex. The C1s signal in Fig. 5(d) shows, besides the component due to the C–C/C–H atoms, a C–O/C–N peak at 285.9 eV and a peak at 289.0 eV, assigned to the carbon of the OC–O–Pd group. The O1s signal in Fig. 5(e) shows a component assigned to the CO/C–O–Pd, a peak assigned to C–OH, and another peak related to physisorbed H2O at 532.0 eV, 533.3 eV, and 535.4 eV respectively.45 Additionally, a peak at 400.8 eV in the N1s signal (Fig. 5(f)) and a peak at 338.8 eV related to the presence of ionic Pd (Fig. 5(d)) are consistent with the predicted structure. Table 1 reports a quantitative analysis of XPS outputs. This data show an N:Pd2+ ratio of about 2:1 and a CO/C–O–Pd:Pd2+ ratio of approximately 4:1, confirming the coordination of Pd2+ with the NH and COO− groups of Pheb molecules.
Element | Atomic concentration |
---|---|
O–Pd | 3.74 |
N–Pd | 4.73 |
Pd2+ | 2.45 |
C1s | 78.06 |
The presence of the Pheb complex in the organic matrix was also confirmed by XPS analysis (Fig. 5(h) and (k)). In the O1s and N1s signals, the peaks at 531.6 eV and 400.8 eV, respectively, indicate the presence of the CO/C–O–Pd and C–NH–Pd bonding in the complex. This assumption is confirmed in the Pd3d peak at 338.4 eV related to ionic palladium, which is related to the Pd–O/N signals. Moreover, a second peak in the N1S and Pd3d peaks related to the CN–Pd and Pd–N at 399.3 eV and 336.8 eV, respectively, can be associated with the bonding between the polymeric matrix and the ionic Pd.
The swelling properties of C-PhebPd were evaluated using a standard gravimetric procedure and compared to a cryogel formed by sola HEMA.46 The newly synthesized macroporous material showed significant swelling capacity. In two seconds, up to 8 times its original weight is achieved (Fig. S12, ESI†). This can be attributed to the synergistic effect of the cryogels’ interconnected porous structure and the pronounced hydrophilicity of the functional groups within the polymer network. Data indicate that palladium complexation did not alter the water uptake capacity of the sponge. This feature is relevant to allowing a fast diffusion of reactant mixture within the catalyst, thus favouring the coupling reaction in a confined area.
Entry | Base | Solvent | Catalyst mol% | Time/min | T (°C) | Yield % |
---|---|---|---|---|---|---|
1 | K2CO3 | H2O/MeOH | — | 30 | 80 | 0 |
2 | K2CO3 | H2O/MeOH | 2.5 | 30 | 80 | 99 |
3 | K2CO3 | H2O/EtOH | 2.5 | 30 | 80 | 99 |
4 | K2CO3 | H2O/MeOH | 2.5 | 30 | 60 | 75 |
5 | Na2CO3 | H2O/MeOH | 2.5 | 30 | 80 | 78 |
6 | K2CO3 | MeOH | 2.5 | 30 | 80 | 15 |
7 | K2CO3 | EtOH | 2.5 | 30 | 80 | 9 |
8 | NaOH | H2O/MeOH | 2.5 | 30 | 80 | 60 |
9 | NaOH | H2O | 2.5 | 30 | 80 | 13 |
10 | TEA | H2O/MeOH | 2.5 | 30 | 80 | 55 |
Further investigations showed that replacing K2CO3 with Na2CO3 (entry 5) led to a slight decrease in product formation. Using organic bases such as TEA (entry 10) caused a significant drop in yields and promoted the formation of undesired side products. Moreover, omitting water and using solely alcohol as the solvent yielded poor product formation, highlighting the importance of solvent composition in this catalytic system. This phenomenon can be elucidated by the swelling behaviour exhibited by our catalyst in aqueous environments. Such swelling facilitates the penetration of reagents into the catalyst's pores, thereby promoting their proximity to each other and enhancing interaction with the catalytic metal species.
The optimal reaction conditions were achieved using K2CO3 as the base and H2O/MeOH (1:1) as the solvent. Under these conditions, the desired product was typically obtained neatly and isolated through a simple extraction process without further purification. All yields were calculated based on the clean, isolated product.
Therefore, the C-PhebPd catalyst application was extended using various aryl halides and boronic acids (Table 3). Starting with standard iodobenzene, we performed the coupling reaction with a range of boronic acids, achieving good to excellent yields for most products (Table 3, entries 1–4, 6–12, and 14). However, the reactions with pyrimidine boronic acid and 2-bromofuran resulted in poor yields (Table 3, entries 5 and 13). In detail, introducing electron-donating substituents on the aryl halide also enhances product yields (Table 3, entries 6–8). Specifically, the synthesis of 4-methoxy-1,1′-biphenyl via the coupling of iodobenzene with (4-methoxyphenyl)boronic acid (Table 3, entry 4) outperforms the reaction between 1-iodo-4-methoxybenzene with phenylboronic acid (Table 3, entry 6), yielding 87% compared to 74%. Similarly, optimal yields for forming 2-phenylfuran were achieved when iodobenzene reacted with furan-2-ylboronic acid (Table 3, entry 2), whereas reversing the reactants (Table 3, entry 13) resulted in minimal product formation. Even in cases where aryls bear electron-withdrawing substituents, the reaction proceeded rapidly, and yields were consistently high (Table 3, entry 14). In most instances, the product quickly precipitated and spontaneously crystallized, simplifying the isolation process and ensuring the purity of the final product. To measure the efficiency of the catalyst, the turnover frequency number (TOF) was calculated for all reactions following eqn (1):
(1) |
Entry | Aryl halides | Arylboronic acid | Time (min) | Productsa | Yieldb (%) | TOF (h−1) |
---|---|---|---|---|---|---|
a Reaction conditions: aryl halide (0.3 mmol), arylboronic acid (0.36 mmol), K2CO3 (0.6 mmol), MeOH/H2O (1:1, 2 mL), C-Pheb-Pd (9 mg, wet), 80 °C. b Yield of isolated product. c Arylboronic acid (2 eq.). | ||||||
1 | 25 | ≥99 | 95.58 | |||
2 | 30 | 89 | 70.27 | |||
3 | 30 | 90 | 71.35 | |||
4 | 25 | 87 | 83.71 | |||
5 | 30 | 14 | 10.81 | |||
6 | 25 | 74 | 62.42 | |||
7 | 20 | 89 | 93.28 | |||
8 | 15 | 58 | 80.21 | |||
9 | 40 | 84 | 55.69 | |||
10 | 30 | 75c | 65.40 | |||
11 | 15 | 45 | 77.83 | |||
12 | 30 | 44 | 32.43 | |||
13 | 120 | Trace | 0.54 | |||
14 | 30 | 94 | 48.08 |
Table 4 compares our catalyst's performance with other heterogeneous catalysts reported in the literature. The table summarizes key metrics such as solvent, time, temperature, and TOF under similar reaction conditions (Table 3, entry 1). This comparison highlights our catalyst's advantages and potential limitations relative to existing alternatives, providing a comprehensive overview of its performance and applicability.
Catalyst | Solvent | Temperature (°C) | Time (h) | TOF (h−1) | Ref. |
---|---|---|---|---|---|
C-PhebPd | H2O/MeOH (1:1) | 80 | 0.41 | 95 | This work |
G-COOH-Pd-10 | H2O/DMF | 70 | 3.00 | 52 | 47 |
PPI-1-NPy-Pd | H2O | 100 | 24.00 | 462 | 48 |
Pd-PEPPSI | H2O/i-PrOH (3:1) | r.t. | 1.00 | 1000 | 49 |
PPI-2-NPy-Pd | H2O | 100 | 24.00 | 685 | 48 |
Pd/MPA | Toluene | 80 | 0.03 | 980 | 50 |
Pd/N-C-300 | n-Butanol | 60 | 24.00 | 198 | 51 |
It is important to highlight that some systems discussed rely on non-sustainable chemicals and demanding conditions, while others utilize more environmentally friendly aromatic catalysts. For example, the catalyst based on Pd immobilized on polyamide (Pd/MPA) demonstrates good catalytic activity at a relatively low cost.50 However, the use of melamine, classified as “carcinogenic to humans”52 and a potential groundwater contaminant, raises serious environmental and health concerns. Moreover, the reaction requires toluene as a solvent and must be performed under a nitrogen atmosphere, limiting its sustainability and safety.
Similarly, while Pd-PEPPSI complexes offer impressive stability in air and water and facilitate reactions at lower temperatures, their synthesis involves harmful solvents like pyridine and dichloromethane, compromising their overall environmental compatibility. Even systems such as PPI-1-NPy-Pd, which allow reactions in water, suffer from lower yields and slower reaction times than our catalyst.49
In contrast, our system addresses these sustainability challenges by avoiding toxic solvents and harmful conditions while delivering high performance with fast reaction times and high yields. This combination of environmental responsibility and catalytic efficiency underscores the potential of our catalyst to drive advancements in green chemistry.
Fig. 6 (a) Recycling test of catalyst in Suzuki–Miyaura reaction between iodobenzene and phenylboronic acid. (b) Yield of standard Suzuki–Miyaura reaction as a function of catalyst contact time. |
A prevalent concern associated with heterogeneous catalysts is the leaching of the metal salt, which poses dual drawbacks: (i) a decline in the catalyst's activity over successive cycles and (ii) potential environmental implications. A test was conducted to confirm the heterogeneity of the C-PhebPd catalyst and assess the potential for metal species leaching. The reaction was initiated and allowed to proceed for a specific duration at 80 °C, varying the contact duration with the catalyst. After removing the catalyst, the reaction continued at 80 °C for 30 minutes. Fig. 6(b) vividly illustrates a decline in catalytic activity upon removing the catalyst from the reaction mixture. These findings unequivocally suggest negligible leaching of palladium ions under the current reaction conditions (Fig. 6(b)).
Mass spectrometry (ICP-MS) was used to further prove the absence of Pd release in water. After the reaction, the Pd content in solution (leaching) was 185 ± 7 ppb.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4qm00800f |
This journal is © the Partner Organisations 2024 |