Ganchang Leia,
Jiayin Wanga,
Xinhui Liub,
Shiping Wang
a,
Shijing Lianga,
Lijuan Shen
*b,
Yingying Zhan
*a and
Lilong Jiang
*a
aNational Engineering Research Center of Chemical Fertilizer Catalyst, State Key Laboratory of Fluorine & Nitrogen Chemicals, School of Chemical Engineering, Fuzhou University, Gongye Road 523, Gulou District, Fuzhou, Fujian 350002, P. R. China. E-mail: jll@fzu.edu.cn; zhanyingying@fzu.edu.cn
bCollege of Environmental and Resource Science, College of Carbon Neutral Modern Industry, Fujian Key Laboratory of Pollution Control & Resource Reuse, Fujian Normal University, 350007, Fuzhou, Fujian, People's Republic of China. E-mail: ljshen@fjnu.edu.cn
First published on 16th June 2025
Developing efficient strategies that convert industrial waste hydrogen sulfide (H2S) into value-added products is meaningful for both applied environmental science and industrial chemistry. Here we report a series of heterogeneous N-doped carbon catalysts with synergistic C–N sites that enable the nucleophilic addition of H2S into aromatic nitrile compounds (PhCN) under mild conditions to produce thiobenzamide (PhCSNH2). The as-designed C–N sites achieve a high thioamide production rate of 26400 μmolPhCSNH2 L−1 h−1 and a notable selectivity of ca. 80% at 60 °C within a short 2-hour timeframe. Additionally, the catalyst exhibits easy recyclability and maintains high stability over ten cycles during a 6-month period. Systematic microscopic and in situ spectroscopic characterization, combined with theoretical calculations, reveal that C-pyridinic N coordination sites effectively lower the adsorption energy barrier of the crucial intermediate *PhCSHNH, offering a dynamically favorable pathway for PhCSNH2 production. Furthermore, the protocol demonstrates excellent compatibility with various substituted substrates, providing access to a diverse range of thioamides.
Broader contextCatalytic conversion of H2S to high value-added organic sulfur-containing compounds is more appealing for the resource utilization of H2S. Herein, we propose an efficient method for synthesizing thiobenzamide (PhCSNH2) through the nucleophilic addition of the cyano group into aromatic nitrile compounds (PhCN) using industrial waste H2S. Heterogeneous N-doped carbon catalysts with synergistic C–N sites were designed for this process. As a result, the C-pyridinic N sites are responsible for the excellent performance of N–GC-X catalysts in PhCSNH2 synthesis under mild reaction conditions, exhibiting a high PhCSNH2 synthesis rate of 26![]() |
From a mechanistic standpoint, the nucleophilic addition of H2S to nitrile compounds primarily involves the dissociation of H2S molecules at active sites, yielding HS* and H*, which further react with the –CN group in nitrile compounds to form thioamides. Research studies have revealed that N-doped carbon catalysts with abundant structural base sites exhibit substantial catalytic activity and remarkable sulfur resistance during the H2S catalytic reaction.19–21 The findings motivate us to undertake the first heterogeneous catalytic synthesis of thioamides from nitriles and H2S. In doing so, we herein fabricate a series of N-doped carbon catalysts with different configurations of N (graphitic N, pyrrolic N and pyridinic N) decorated on two-dimensional porous carbon materials. The catalyst with rich pyridinic N exhibits an optimal PhCSNH2 yield of up to approximately 26400 μmolPhCSNH2 mL−1 h−1 with good product selectivity (>80%) and high durability through 10 runs of cyclic testing over 6 months. A wide variety of substituted substrates are viable in this reaction, enabling access to diverse thioamides. Using detailed in situ spectroscopic characterization, combined with theoretical calculations, the common mechanistic pathway is elucidated. The progress is anticipated to inspire more research attention on the cost-efficient synthesis of thioamides with high atom economy, as well as open the way for highly efficient resource utilization of the H2S waste.
Fig. S2 (ESI†) shows the SEM images of the as-prepared samples. Pristine GC presents a layered block-like structure (Fig. S2a, ESI†). As the dicyandiamide content increases, the obtained N–GC-X catalysts gradually decrease in size, showing thin-layer morphologies with holey structures rather than stacked layers (Fig. S2b–e, ESI†). Further insight from TEM characterization highlights the unique tulle-like morphologies of the N–GC-6 catalysts (Fig. 2a and b). EDS mapping demonstrates uniform distribution of carbon, nitrogen, and oxygen elements in the layered structure (Fig. S3, ESI†). In addition, the TEM analysis uncovers the presence of mesopores, revealing the porous nature of N–GC-6. The layered structure of N–GC-6 with abundant nanoscale pores is also corroborated by the AFM image, as illustrated in Fig. 2c. The randomly measured porous nanosheets have a similar thickness of about 1.8–2.3 nm. Furthermore, Fig. S4, S5 and Table S1 (ESI†) present the nitrogen adsorption–desorption isotherms and the corresponding BJH pore size distributions for the N–GC-X samples. All N–GC-X catalysts exhibit type IV isotherms with narrow H2 hysteresis loops, confirming mesoporous properties according to the IUPAC definition.
To obtain more microstructural information of the catalysts, multiple spectroscopic characterization studies have been conducted. Fig. S6 (ESI†) displays the XRD patterns of the as-prepared samples, which all show two characteristic peaks at around 26° and 44° that are assigned to the (002) and (101) crystal planes of graphitic carbon, respectively.22,23 The finding indicates the formation of amorphous carbon networks in the N–GC-X catalysts. Additionally, N–GC-X exhibits a relatively weak peak intensity and a high angular shift in the (002) plane, indicating the presence of numerous structural defects in its amorphous carbon network. The chemical functional groups in the samples have been investigated by FT-IR characterization (Fig. S7, ESI†). The N–GC-X samples exhibit three bands at around 1320–1197 cm−1, 1608–1515 cm−1, and 2200 cm−1, corresponding to the CN, C–N, and O
C
N groups, respectively.24 This result confirms the successful introduction of functional N species into the N–GC-X framework.
Moreover, in the Raman spectra, both pristine GC and N–GC-X display peaks at around 1350 cm−1 and 1590 cm−1, which are attributable to the disorder-induced D-band and the in-plane vibrational G-band, respectively (Fig. S8, ESI†). Notably, the intensity ratio of the D-band to the G-band (ID/IG) increases from 0.98 to 1.07 for the N–GC-X samples, indicating an increased graphitization degree and elevated topological defects.25,26 This deduction is consistent with the XRD analysis results, implying that the incorporation of N species benefits the formation of structural base sites. Fig. S9 (ESI†) displays the 13C NMR spectrum of N–GC-6, which is obtained to unveil the C and N chemical bonding structure of the sample. The bands at 98 ppm and 129 ppm are attributed to aromatic carbon in the pyridine ring. The band at 170 ppm could be attributed to carbon substituted by the CN cross-linker in the pyridine ring.27–29 The results agree well with the results of the following XPS analysis, that is, the doped N in the N–GC-X catalyst is mainly pyridinic N. As depicted in Fig. S10 (ESI†), the C 1s XPS spectra of all N–GC-X catalysts can be deconvoluted into three peaks at around 284.6, 285.9, and 287.8 eV, attributable to the CC/C–C, C–N, and N
C signals, respectively.30 In contrast, pristine GC displays only one predominant peak located at approximately 284.6 eV, corresponding to the C
C/C–C unit. The N 1s spectra in Fig. 2d–g and Fig. S11 (ESI†) are fitted into four peaks with binding energies of 398.6 eV, 399.7 eV, 400.7 eV and 401.8 eV, which are assigned to pyridinic N, pyrrolic N, graphitic N, and quaternary N, respectively.31,32 The above findings indicate that the structural N sites are successfully doped into the carbon skeleton. Fig. 2h depicts the total N content and the distribution of N species in the catalysts. It can be found that the total N content in the samples follows the order N–GC-6 (39.7 at%) > N–GC-8 (39.3 at%) > N–GC-4 (33.9 at%) > N–GC-2 (26.8 at%). N–GC-6 possesses the highest N content. The results are also validated by the elemental analysis (EA) (Table S1, ESI†). Notably, with the variation of the dicyandiamide content, the percentage of pyridinic N in the series of N–GC-X samples changes from 43.6% to 48.0%, surpassing that of pyrrolic N (from 24.8% to 26.4%) and graphitic N (from 27.2% to 30.0%) (Fig. 2i). The different N species in all N–GC-X samples follow the order pyridinic N > pyrrolic N ≈ graphitic N > quaternary N. It should be noted that quaternary N usually appears inert in heterogeneous catalytic reactions. Thus, we mainly investigate the most adequate pyridinic N, pyrrolic N and graphitic N in the present work. The percentages of pyridinic N in all N–GC-X samples are higher than 43%, which plays a key role in the H2S addition reaction (as discussed later). Collectively, the above characterization methods faithfully validate that we have successfully synthesized a series of N–GC-X catalysts with different weight contents of N doping by varying the amounts of precursors. Importantly, the distribution of N species in these samples has also been modulated, which establishes a foundational framework for identifying active sites and conducting in-depth mechanism studies.
To further elucidate the relationship between the structure of the catalysts and their catalytic properties, kinetic experiments have been performed at low conversions of <20% to achieve accurate kinetic parameters. The ordinary rate formula for the nucleophilic addition of PhCN with H2S can be derived using the following eqn (1). Since the amount of catalyst is constant and the concentration of H2S is significantly higher than the PhCN concentration, eqn (1) can be simplified to eqn (2). As shown in Fig. 3c and d, the curves that depict the decrease in the concentration of PhCN (CPhCN) as a function of reaction time (t) give a straight line from the starting point, denoting a pseudo-zero-order reaction (γ = 0) with respect to PhCN. Thus, eqn (2) can be further expressed as eqn (3) or (4). The experimental results suggest that the reaction rate is in agreement with eqn (4), and there is a linear relationship between CPhCN and t. Based on the Arrhenius equation: k = A e−Ea/RT, the apparent activation energy Ea can be calculated using eqn (5). The PhCSNH2 production rate over N–GC-6 displays an approximately linear enhancement with increasing temperature, ranging from 8600 to 26400 μmolPhCSNH2 L−1 h−1 when the reaction temperature increases from 40 to 70 °C (Fig. 3e). The results indicate that the reaction rate of catalysts is enhanced with the incorporation of structural N base sites. As depicted in Fig. 3f, the extraction of the activation energy Ea from the Arrhenius curve leads to Ea = 35.8 kJ mol−1 and 36.4 kJ mol−1 for N–GC-6 and pristine GC, respectively, in line with the observed trend in catalytic activity. The findings prove that the synergistic effect between C and N sites can significantly lower the Ea and enhance the catalytic performance. By comparing the relationship between the pyridinic N content and the yield of PhCSNH2 for N–GC-X materials (Fig. 3g), it is obvious that the increase of pyridinic N content of the catalysts enhances the PhCSNH2 yield. Furthermore, the optimized N–GC-6 displays better reactivity than the traditional metal-based catalysts, including Fe2O3, CoO, CeO2 and CaO (Fig. S15, ESI†). More importantly, N–GC-6 exhibits stable activity in H2S nucleophilic addition, with no obvious decline of PhCSNH2 selectivity even with the introduction of 1% CO2 (Fig. S16, ESI†). The above test findings clearly reveal that N–GC-6 exhibits high activity and outstanding PhCSNH2 selectivity, which is positively correlated with its high pyridinic N content.
![]() | (1) |
![]() | (2) |
![]() | (3) |
CBN = −k·t | (4) |
ln![]() ![]() | (5) |
Given the high catalytic performance, the catalytic stability of N–GC is tested as it is crucial for practical application. The N–GC-6 catalysts were reused for five cycles via facile centrifugation separation without any other treatment. As exhibited in Fig. 3h, there is no significant decrease in both activity and PhCSNH2 selectivity during the cycle test. N–GC-6 achieves a PhCN conversion of 97% with a PhCSNH2 selectivity of 78% after five runs of the reaction. In addition, after storing under ambient conditions for 6 months, N–GC-6 maintains a high catalytic activity and structural stability comparable to those of a fresh catalyst (Fig. 3h and Fig. S17–S20, ESI†), demonstrating its excellent stability. Besides, the slight decrease in reactivity over the catalysts may be caused by the generation of sulfur species that can cover the active sites or block the pores of the catalysts.
With the confirmed optimal reaction parameters, we then explored the reaction scope for producing different thioamides by nucleophilic addition of the cyano group into nitrile compounds using H2S over an N–GC-6 catalyst. As displayed in Table 1 and Fig. S21 (ESI†), the reaction proceeds smoothly in all cases, no matter for strongly electron-withdrawing or electron-donating substrates. For instance, following a reaction time of 2 h at 60 °C, p-chlorobenzonitrile exhibits a 92% conversion and a good regioselectivity of 80% for the p-chlorothiobenzamide product. Substrates bearing a hydroxyl group also display nearly 80% selectivity for the generation of 4-hydroxythiobenzamide. In brief, N–GC-6 exhibits excellent tolerance to functional groups including chlorine, bromine, fluorine, and methyl. Additionally, the approach is found to be effective in the production of aliphatic thioamides with high yields using acetonitrile and 2-methylpropanenitrile as substrates (thioacetamide and 2-methylpropanethioamide). Furthermore, even upon switching the reaction solvent to DMF, N–GC-6 still demonstrates comparable catalytic performance in transforming the nitrile-based compounds into their corresponding thioamides, highlighting the unique role of the pyridinic N sites in promoting nucleophilic addition reactions.
The calculation results reveal that the PhCN molecule can be intensely adsorbed on the C-pyridinic N site and facilely activated to *PhCN with a low activation energy of −0.49 eV. In contrast, adsorption of the PhCN molecule on the C-graphitic N site and the C-pyrrolic N site is less favorable, requiring higher activation energies of −0.21 eV and −0.32 eV, respectively (Fig. 4b and Fig. S22–S24, ESI†). Additionally, the dissociation of *H2S into *HS and *H on the N sites is found to be endothermic, with free energies of −0.41 eV, −0.62 eV and −0.81 eV for graphitic N, pyrrolic N and pyridinic N, respectively (Fig. 4b). A transition barrier energy (TS1) of 0.91 eV is required for the pyridinic N site, significantly lower than the values of 1.54 eV for graphitic N and 1.23 eV for pyrrolic N (Fig. 4c). The higher dissociation energies and TS1 energies of the graphitic N and pyrrolic N sites suggest that the dissociation of H2S on these sites is more challenging, which would result in relatively low activity.
Following dissociation (Fig. 4c), the *H reacts with *PhCN to form the *PhCNH intermediate, with TS2 values of 0.60 eV, 0.44 eV and 0.19 eV for graphitic N, pyrrolic N and pyridinic N, respectively. The formation energies of *PhCNH on graphitic N and pyrrolic N are −0.62 eV and −1.03 eV, compared to −1.25 eV on the pyridinic N, indicating the high activity of the pyridinic N site for the H2S nucleophilic addition reaction. Subsequently, *HS reacts with *PhCNH to form the key intermediate *PhCSHNH, with a TS3 of 0.85 eV for pyridinic N. This value is notably lower than those of graphitic N (1.75 eV) and pyrrolic N (1.18 eV). Importantly, the pyridinic N site can facilitate the –HS spillover on the catalyst surface, leading to the formation of *PhCSHNH species with a free energy of −1.60 eV. In contrast, graphitic N and pyrrolic N exhibit higher formation energies. The *PhCSHNH intermediate then undergoes intramolecular rearrangement of the –SH group to further convert into *PhCSNH2, with a TS4 of 0.56 eV for pyridinic N. Furthermore, the formation energy barrier for PhCSNH2 formation on pyridinic N (−1.56 eV) is much lower compared to that on graphitic N (−1.32 eV) and pyrrolic N (−1.40 eV), indicating that the PhCSNH2 product is more readily generated at the pyridinic N site. Additionally, the PhCSNH2 desorption free energies at the N species follow the order pyridinic N > graphitic N > pyrrolic N, aligning with the activity trend (Fig. 4d). These findings suggest that the pyridinic N centers can serve as efficient sites for the H2S nucleophilic addition. According to the calculations, the rate-determining step of the H2S nucleophilic addition reaction is the nucleophilic attack of *PhCNH by *HS to form *PhCSHNH (Fig. S25a, ESI†). The lower activation energies for the formation of the *PhCSHNH intermediate on pyridinic N consolidate that a higher proportion of pyridinic N facilitates the nucleophilic reaction (Fig. 4e). Moreover, the *PhCSHNH key intermediate serves as a probe to determine the formation enthalpies of the C–S bond (ΔHC–S) on different N sites. The ΔHC–S is calculated to be −33.6 kJ mol−1 on pyridinic N. In contrast, on pyrrolic N and graphitic N, the ΔHC–S is substantially enhanced to −0.29 kJ mol−1 and −0.58 kJ mol−1, respectively. The lower ΔHC–S on pyridinic N signifies that the formation of the C–S bond to generate *PhCSHNH becomes more beneficial, leading to a markedly distinct selectivity in the production of PhCSNH2 on pyridinic N.
The charge density difference and Bader charge analyses were also performed to evaluate the affinity of N sites over reactive atoms (Fig. S25b, ESI†). The variation in charge difference among the three N site-PhCSHNH intermediates is 0.48–0.57 eV, verifying the hypothesis that introducing N atoms enhances the affinity of C sites on the carbon matrix by creating basic sites. The results are consistent with the CO2-TPD characterization results that the reactants are strongly adsorbed on the surface of N–GC-6 due to its plentiful structural base sites (Fig. S26, ESI†). In addition, the pyridinic N sites show weak interaction with PhCSNH2, and the charge shift from the pyridinic N site to PhCSNH2 is 0.45 eV, slightly exceeding the values for graphitic N (0.39 eV) and pyrrolic N (0.35 eV). The comprehensive theoretical calculation suggests that *PhCSHNH is favorable for addition rearrangement to *PhCSNH2 at the pyridinic N-based basic nitrogen sites.
To further explore the reaction process of nucleophilic addition of PhCN with H2S on the catalyst, Quasi-in situ IR spectroscopy of PhCN and H2S co-adsorption were conducted to investigate the real-time intermediates of the reaction (Fig. 5a). The recorded IR spectra display a characteristic band at around 2232 cm−1, which is assigned to the CN group.33 Concurrently, a new band at around 1390 cm−1 assigned to ν(H–S) is detected.34 The intensity of the C
N group gradually weakens with reaction time, while two vibration bands emerge at around 1758 cm−1 and 1630 cm−1, which can be attributed to the ν(C
N) and δ(N–H) groups, respectively.35,36 Notably, the bands attributable to the C
S and C
S–H groups are detected at around 2188 cm−1 and 990–1210 cm−1, respectively, confirming the generation of *PhCSHNH as a key intermediate in the H2S nucleophilic addition reaction.37,38 This observation aligns with the DFT calculation results. Furthermore, the intensities of the N–H and C
S groups enhance with time, revealing the formation of Ph-CSNH2 during the reaction.
Quasi-in situ Raman spectra of H2S and PhCN co-adsorption on N–GC-X have also been recorded to shed light on the reaction process (Fig. 5b and c). The characterization was conducted under experimental conditions, i.e., atmospheric pressure, T = 60 °C, using 10 mL PhCN + 5 wt% H2S (balance N2). Before evaluation, 10 mL of PhCN and 40 mg of the catalyst were carefully added into a reaction tank, followed by the continuous introduction of 5 wt% H2S into the reaction system. It is observed that the intensity of the characteristic peak at 2230 cm−1 associated with ν(CN) is decreased with reaction time,39 while the intensity of the –NH2 group (3195 cm−1) is enhanced. The phenomenon indicates that H atoms in H2S are easily added to PhCN to form NH/NH2 groups, demonstrating a pronounced nucleophilic addition of H2S. Additionally, a band at 2576 cm−1 attributed to the generation of the *HS intermediate is detected,40 further validating the nucleophilic addition between H2S and PhCN.
Furthermore, four bands located at 377 cm−1, 459 cm−1, 550 cm−1 and 1490 cm−1 can be attributed to the ν(CS) and C–S unit vibrations,41,42 respectively. The intensity of these peaks gradually enhances with reaction time, indicating the enrichment of CS species on the catalyst surface. The above results suggest that PhCN is first reduced to *PhCNH by dissociated *H in H2S. *PhCNH is then attacked by *HS to generate *PhCSNHNH, which finally rearranges into *PhCSNH2. The results of Quasi-in situ IR and Quasi-in situ Raman characterization are consistent with the calculated activation barriers for the formation of *PhCNH, *PhCSHNH and *PhCSNH2 intermediates on the pyridinic N (Fig. 5d), which reveals that *PhCSHNH is the rate-limiting intermediate in this reaction. Overall, the formation of structural base sites is the main contributor to high activity and product selectivity in the H2S nucleophilic addition reaction (Fig. 5e).
For the reusability evaluation, the catalysts were separated by centrifugation after the end of the reaction and washed with ethanol three times, followed by drying in vacuo before the next catalytic cycle.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ey00110b |
This journal is © The Royal Society of Chemistry 2025 |