Siwei Yang‡
*ab,
Caichao Ye‡cd,
Xun Songa,
Lin Hea and
Fang Liao*a
aChemical Synthesis and Pollution Control Key Laboratory of Sichuan Province, College of Chemistry and Chemical Engineering, China West Normal University, Nanchong 637002, P. R. China. E-mail: liaozhang2003@163.com; Fax: +86 817 2568067; Tel: +86 817 2568067
bState Key Laboratory of Functional Materials for Informatics, Shanghai Institute of Microsystem and Information Technology, Chinese Academy of Science, Shanghai 20050, P. R. China. E-mail: yangsiwei@mail.sim.ac.cn; Fax: +86 021 62511070 ext. 420; Tel: +86 021 62511070 ext. 420
cSchool of Chemical Engineering, Nanjing University of Science and Technology, Nanjing 210094, P. R. China
dMaterials and Process Simulation Center, California Institute of Technology, Pasadena 91125, USA
First published on 17th October 2014
A magnetically recyclable and highly photocatalytic poly(p-phenylenediamine)–Fe3O4 (PpPD–Fe3O4) composite photocatalyst was synthesised by a one-step chemical oxidation polymerisation based on theoretical calculation results. The band gap and selectivity of PpPD for acid dyes were studied by theoretical calculations. The calculation results showed that the PpPD molecules exhibited better conjugacy than PANI and that the band gap was more fit for use as a photocatalyst. The photocatalytic activities were evaluated by the degradation of various dyes. The stability and high selectivity of this composite was remarkable. The combination of Fe3O4 with PpPD lead to high photocatalytic activity in the degradation of the dyes under both ultraviolet, and visible light, irradiation, although Fe3O4 alone was inactive as a visible-light-driven photocatalyst.
As a new type of photocatalyst, conducting polymers (CP), such as polyaniline (PANI), poly-pyrrole, polythiophene, and their derivatives or composites have received increasing attention due to their advantages of low cost, ease of synthesis, high conductivity, and environmental stability.18–30 Wang et al.1 have synthesised a CuI/PANI super-hydrophobicity photocatalyst. Xiong et al.31 found a highly enhanced performance CoFe2O4–PANI magnetically recyclable photocatalyst. The combination of CoFe2O4 nanoparticles with PANI lead to a high photocatalytic activity in the degradation of dyes under visible light irradiation and the catalyst showed better catalytic activity for anionic dyes. However, how to improve the stability and catalytic activity of CP-based photocatalysts remains a challenge to researchers. Furthermore, the design of a CP-based photocatalyst which is easy to retrieve, or with high selectivity, is also a worthwhile research topic.
In this work, an Fe3O4–poly(p-phenylenediamine) (Fe3O4–PpPD) composite was prepared by an in situ polymerisation of the p-phenylenediamine (pPD) monomer in aqueous solution based on theoretical calculation results. All the theoretically calculated results were in good agreement with the experimental outcomes. Various dyes were used to evaluate the catalytic activity of the Fe3O4–PpPD composite. This composite can be separated easily by an external magnetic field and showed high selectivity and catalytic activity for acid dyes under both UV and visible light irradiation. The calculated results were consistent with the experiments. Fe3O4 which have no catalytic activity served as an electron sink and prevented the recombination of photoinduced electrons and holes. The PpPD is excited by UV or visible light and obtained photogenerated carriers and excited electrons, respectively. The excited state electrons in PpPD can readily migrate to the conduction band of Fe3O4. As PpPD is a good material for transporting holes, the photogenerated charges can migrate easily to the surface of the photocatalysts and photodegrade the adsorbed dye molecules.
For the case of the adsorption configurations, the corresponding bond dissociation energy (EBDE) was calculated according to the expression
EBDE = E(P–D) − (EP˙ + ED˙) | (1) |
The control sample (PpPD) was prepared as follows: 5.0 mL (10 mM) APS and 2.0 mL (1 M) HCl were mixed in a beaker. Then, 5.0 mL (10 mM) pPD were added to this solution. Shaking stopped after 0.5 min, and the reaction solution was kept for 24 h at 25 °C. The black product was collected by filtration, washed twice with distilled water and then dried in vacuum.
Previous reports confirmed that the PANI and PpPD were mixtures of oligomers, include dimer, trimer, and other oligomers.24–26,36,37 Considering this issue, four types of oligomers of PpPD were chosen: Fig. 1 shows all types of molecular structures used in the calculations, A1 was the molecular structure of the PANI, A2, A3, A4, and A5 were the molecular structures of PpPD. B1 to B4, and C1 to C4 were molecular models of PpPD, with which the BB combined by chemical adsorption. B1 to B4 were the BB combined with the N atoms (N–) which were in the phenazine skeleton of PpPD. C1 to C4 were the BB combined with the N atoms (–NH2,
NH) which were in the end-groups and residues of PpPD. D1 and D2 were the molecular models of PpPD, with which the MO combined at different sites (both the N atoms in the phenazine skeleton or end-groups) of PpPD.
Fig. 2 shows the molecular configurations and electric cloud distribution of the HOMO–LUMO forms of PANI and PpPD, respectively. There was an angle between the planes of the adjacent benzene rings in PANI. However, all the atoms of the PpPD were mainly located in same plane, this suggested that the planarity of PpPD was better than PANI (Fig. 3). The phenazine skeleton25 limited the rotation of the benzene ring which improved the rigidity of PpPD. Additionally, the π electron cloud was delocalised over the whole molecular plane of the PpPD, while the π electron cloud only delocalised in the separation benzene rings of PANI. This indicated that PpPD molecules exhibited better conjugacy than PANI, and the carrier mobility of PpPD was also higher than that of PANI.
![]() | ||
Fig. 2 Ball-and-stick model and theoretical electron distribution of the HOMO–LUMO energy states of PANI (A1) and PpPD (A2–5). |
The band gap (BG, eV) and the energy levels of molecular orbitals (HOMO, LUMO) are listed in Table 1. The BG of different types of PpPD polymer were 0.58 to 1.85 eV, this indicated that the PpPD acted as a type of semiconductor like PANI. Additionally the BG of PpPD was slightly lower than that of PANI,31 which showed that the PpPD might have well catalytic activity under visible light irradiation.
Molecule | GB (eV) | HOMO (eV) | LUMO (eV) |
---|---|---|---|
A1 | 1.18745 | −5.08965 | −3.90220 |
A2 | 1.85440 | −4.63269 | −2.77828 |
A3 | 1.08584 | −4.68689 | −3.60105 |
A4 | 2.05174 | −4.96839 | −2.91665 |
A5 | 0.58921 | −4.68491 | −4.09570 |
Fig. 4 showed the molecular configuration of B1–B4, C1–C4, D1, and D2. The calculated results showed that both acid and alkaline dyes could be chemisorbed on PpPD. Table 2 showed the bond length (Å), and bond dissociation energy between PpPD and the dyes. The bond lengths were 1.4 to 1.5 Å when the BB or MO combined with the N atoms in both phenazine skeleton and end-groups of PpPD. This indicated that PpPD had strongly interacted with BB or MO. The short bond between BB and PpPD could be attributed to the strong interaction between acid groups (phenolic hydroxyl groups of BB) of acid dyes and alkalinity groups of PpPD. It was interesting that, although the bond lengths between dyes and PpPD were almost identical, the bond dissociation energy showed significant differences when the dyes combined with the N atoms at different loci of PpPD. The bond energy was lower than 100 kJ mol−1 when the dyes combined with the N atoms in the phenazine skeleton of PpPD.
Molecule | Bond length (Å) | Bond dissociation energy (kJ mol−1) | BG (eV) |
---|---|---|---|
B1 | 1.418 | 10.0 | 1.44756 |
B2 | 1.407 | 68.6 | 1.44521 |
B3 | 1.442 | 6.30 | 1.42303 |
B4 | 1.417 | 91.1 | 0.92779 |
C1 | 1.441 | 179.8 | 1.67881 |
C2 | 1.535 | 159.8 | 1.11635 |
C3 | 1.447 | 201.0 | 2.05490 |
C4 | 1.459 | 214.7 | 0.91419 |
D1 | 1.414 | 24.5 | 0.37397 |
D2 | 1.407 | 298.1 | 0.59422 |
Moreover, the bond energy increased (150 to 300 kJ mol−1) when the dyes combined with the N atoms in the end-groups and with the residues of PpPD. This change in bond energy could be attributed to the differences in N atoms and their loci on the PpPD. The large sterically hindrance of the N atoms in the phenazine skeleton mades the structure unstable and easy to desorbed when the dyes were adsorbed on PpPD. On the other hand, the amino and imino groups located in the end-groups and residues of PpPD, the steric effect was much weaker and the structure was more thermodynamically stable. There are many amino and imino groups in PpPD which ensured the high adsorption ability of the catalysts because of their low degree of polymerisation.31
All of the results demonstrated that the PpPD exhibited no selectivity for acid or alkaline dyes in the absorption progress. However, electrostatic repulsion might have affected the adsorption process when the alkaline dyes were closer to the N atoms of PpPD. Additionally, the selectivity in catalytic progression must be considered. The BGs of PpPD which adsorbed the dyes are also shown in Table 2. The BGs of PpPD changed after the dyes were adsorbed on the PpPD. The BG varied over a small range (0.9 to 2.0 eV) when the bromophenol blue combined with the N atoms in the phenazine skeleton or end-groups of the PpPD. Nevertheless, the BG decreased to 0.3 to 0.5 eV when the alkaline dye was adsorbed on the PpPD. The BG decreased: this may have resulted because the superoxide radical and hydroxyl free radical were harder to generate due to the low energy of the hole and electron. This may have caused the PpPD to degrade the alkaline dyes less readily.
Therefore, the PpPD might have selectivity to acid dyes in the photodegradation process. The alkaline dye could not easily remain in close proximity to the N atoms of PpPD due to the innate electrostatic repulsion in the reaction kinetics, although it was stable when the alkaline dye combined with the N atoms in both phenazine skeleton and the endgroups of PpPD. Additionally, the strong interaction between an alkaline dye and the PpPD caused the BG to decrease significantly which could result in a decrease in degradation ability towards alkaline dye. So the PpPD might be a selective photocatalyst in both adsorption and catalytic processes.
The Raman spectra were used to analyse the molecular structure of the PpPD–Fe3O4 composite (Fig. 6a). The peak at 1610 cm−1 was ascribed to the benzene ring. The peak at 1517 cm−1 corresponded to the N–H bending deformation mode and the bands at 1549 and 1594 cm−1 were ascribed to the C–C deformation of benzenoid rings and quinoid rings [35, 36], respectively. The spectra also revealed that the peaks located between 1361 and 1430 cm−1 were assigned to the C–N+ stretching modes of the delocalised polaronic charge carriers. The peak at 1164 cm−1 was due to the C–H inplane bending mode, which indicated the presence of a quinoid ring. The weak band at 1199 cm−1 was ascribed to the C–N stretching mode of polaronic units.21–24
The FT-IR spectrum of PpPD–Fe3O4 composite is shown in Fig. S4, ESI.† The adsorption peaks at 3321, 3240, and 3210 cm−1 corresponded to the N–H stretching mode, and implied the presence of secondary amino groups. The strong peaks at 1601 and 1536 cm−1 were ascribed to CN and C
C stretching vibrations in the phenazine structure, respectively. The peaks at 1239 and 1370 cm−1 were associated with C–N stretching in the benzenoid and quinoid imine units.37–40 The peaks at 900 and 830 cm−1 can be attributed to the out-of-plane deformation of C–H on a 1, 2, 4-tetrasubstituted benzene ring, which implied that the polymers had a basic phenazine skeleton. Moreover, the bands at 754 and 593 cm−1, characteristic of C–H out-of-plane bending vibrations of benzene nuclei in the phenazine skeleton, were also observed.41–44 The MALDI-TOF-MASS spectrum of this PpPD composite (Fig. S5, ESI†) confirmed that the PpPD was an oligomer and mainly composed of tripolymer, tetramer, and pentamer.
Fig. 6b showed the XRD patterns of the PpPD–Fe3O4 composite, pure Fe3O4, and PpPD. The two broad diffraction peaks at 20 and 30° were attributed to the amorphous structure of PpPD.37 Each diffraction peak of the Fe3O4 could be assigned to the cubic phase structure of Fe3O4 (JCPDS 65-3107). The peaks at 2θ values: 30, 35, 43, and 62.5°, could be indexed as the (220), (311), (400), and (422) crystal planes of cubic phase Fe3O4, respectively. There were no obvious diffraction peaks of PpPD in this composite which was similar to amorphous PANI in composites based on PANI with simple metal oxides.39
Fig. 6c showed the UV-vis spectra of the PpPD–Fe3O4 composite. The UV-vis absorption spectrum in H2O revealed absorption maxima at 545 and 350 nm, originating from the charge-transfer-excitation-like transition from the highest occupied energy level to the lowest unoccupied energy level and the π–π* transition, respectively.31,34 The results showed good agreement with the theoretical calculation and implied that the PpPD–Fe3O4 composite can be used as a UV or visible-light-driven photocatalyst. The N2 adsorption–desorption isotherm and corresponding pore size distribution of the PpPD–Fe3O4 composite are shown in Fig. 6d. This composite showed Type III isotherms, which exhibited almost no hysteresis loops. It indicated that N2 was hardly absorbed on the surface of the PpPD–Fe3O4 composite, and the phenomenon of autoacceleration was exhibited as the adsorption process occurred. The corresponding pore size distribution, as calculated by the BJH method, is shown in the insert to Fig. 6d: the pore size distribution of the PpPD–Fe3O4 composite consisted of small particle pores (2 to 12 nm), where the average pore diameter was 2 to 50 nm. The BET surface areas, total pore volumes, and average pore diameters of PpPD–Fe3O4 composite are summarised in Table. S1, ESI.† The decomposed XPS Fe 2p spectra of PpPD–Fe3O4 are shown in Fig. S7.†
To estimate the selectivity to acid dyes of PpPD–Fe3O4, Fig. 9a and b show the photocatalytic degradation of BG, RhB, MB, CR, and MO under UV and visible light in the presence of PpPD–Fe3O4. Fig 9c and d show the rate constants for various dyes (BG, BB, BP, RhB, NR, MB, SIII, MO, and CR). The photodegradation process could be expressed by pseudo-first-order kinetics:
![]() | (2) |
All the dyes degraded completely in 24 h in the presence of PpPD–Fe3O4 (Fig. S8–11, ESI†). This showed that the PpPD–Fe3O4 displayed good catalytic activities on the photodegradation of acid dyes (BG, BB, BP, RhB, NR, MB, and SIII) under both UV and visible light irradiation. Furthermore, the alkaline dyes (CR and MO) could not be degraded promptly by PpPD–Fe3O4. The photocatalytic degradations (Table 3) of the dyes were (in order): RB > BG > BP > SIII > BB > NB > MB ≫ CR > MO. The selectivity of the catalysts was significant and the results were in good agreement with the theoretical calculations.
Dyes | UV light | Visible light | ||
---|---|---|---|---|
C/C0 | K (h−1) | C/C0 | K (h−1) | |
BG | 0.00941 | 0.25922 | 0.08758 | 0.13529 |
BB | 0.05472 | 0.16142 | 0.18233 | 0.09455 |
BP | 0.02416 | 0.20684 | 0.12256 | 0.11662 |
RB | 0.00164 | 0.35628 | 0.09521 | 0.13065 |
NB | 0.09076 | 0.13331 | 0.13468 | 0.11138 |
MB | 0.11076 | 0.12224 | 0.21893 | 0.08439 |
SIII | 0.03076 | 0.19432 | 0.15173 | 0.10476 |
MO | 0.82051 | 0.01099 | 0.82051 | 0.01099 |
CR | 0.85557 | 0.00867 | 0.85116 | 0.00895 |
The photocatalytic activity of the recycled PpPD–Fe3O4 was also investigated, and the results are shown in Fig 9e and f. There was slight change in the photocatalytic activity of the catalyst in the first 10 cycles under visible or UV light irradiation. Further experimental data (Fig. S12 and S13, ESI†) verified the fact that the photocatalytic activity reduced significantly after 25 and 15 cycles under visible and UV light irradiation respectively. The lower stability of the catalyst could be attributed to the destructive effect of UV light on the PpPD. These results indicated that the PpPD–Fe3O4 catalyst showed good stability under both visible and UV light irradiation and was better than PANI.12–19 Furthermore the magnetism of the catalyst made it easy to recycle.
As shown in Fig. 10, both ultraviolet light and visible light irradiation were restrained, after the injection of EDTA-2Na, with regard to the photodegradation of BB. Futhermore, the photo-degradation ratio of BB was only slightly changed by the addition of t-BuOH under both ultraviolet light and visible light irradiation. This phenomenon suggested that holes were the main active species in the development and progress of photocatalysis.45 Furthermore, the photocatalytic activity was enhanced in the presence of Fe3O4. The separation efficiency of photoinduced electrons and holes was enhanced by the electronic interaction between Fe3O4 and PpPD.12–14
To further verify the selectivity of the catalyst, as predicted by theoretical calculation and confirmed by photocatalytic degradation experiment, the adsorption kinetics and isotherm of the BB on PpPD–Fe3O4 catalysts were studied. Kinetics experiments were carried out to verify whether or not the system followed the Langmuir–Hinshelwood mechanism. The equilibrium uptake could be calculated as follows:
![]() | (3) |
Fig. 11 shows the adsorption capacity of BB and MO over time in the presence of the PpPD–Fe3O4 composite: there was significant adsorption in the case of the BB. This was due to the acid groups of the BB which underwent chemical interactions with the alkalinity groups of PpPD. In contrast, the alkalinity dyes with alkalinity groups could not easily gain access to the alkalinity groups of PpPD because of electrostatic repulsion, eventually resulting in poor adsorption of the dyes. These experimentally measured results agreed with those calculated theoretically.
The rate kinetics of BB adsorption on the PpPD–Fe3O4 catalysts were analysed using the 1st-order rate kinetic model proposed by Lagergren (2),46 a pseudo-2nd-order model (3),47 and the rate equation as expressed by Weber and Morris (4):48
![]() | (4) |
![]() | (5) |
![]() | (6) |
Fig. 11b shows the Lagergren plot of log(qe − q) versus t (min) for adsorption of BB by the PpPD–Fe3O4 composite. The linearity of this plot indicated that a 1st-order mechanism did not follow the pseudo-1st-order kinetic model. The pseudo-2nd-order equation was expressed as a linear plot of (t/q) versus t (Fig. 11c) and was used to determine the rate constants and correlation coefficients. The pseudo-2nd-order rate constant (K′) and correlation coefficient (r2) for removal of BB from an aqueous solutions of PpPD–Fe3O4 composite were 0.00202 mg g−1 min−1 and 0.9991 respectively. This suggested that the values of correlation coefficients indicated a good fit for the pseudo-2nd-order model with the experimental data.49–51 Fig. 11d showed the Weber and Morris (intra-particle diffusion) plot for the adsorption of BB by PpPD–Fe3O4 composite: the plot was not linear over the entire time range. The kinetics experiments indicated that the mechanism of BB adsorption was complex and both the surface adsorption as well as intra-particle diffusion contributed to the rate determining step.48 This showed that the adsorption was mainly chemical adsorption51 and the calculation results agreed with the experimental measurements.
The BB adsorption isotherm data were correlated with the theoretical models of Langmuir52 and Freundlich:53
![]() | (7) |
![]() | (8) |
The linear plot of Ce/qe versus Ce and was used to determine the value of Langmuir constants qmax (mmol g−1) and b (L mg−1) (Fig. 11e). The Langmuir constants qmax and b were 505.05 mg g−1 and 9.68 × 103 L mg−1, respectively, furthermore, the correlation coefficient (r2) was 0.9973. The r2 values indicated a good fit to the Langmuir model for the experimental data.52,53 Fig. 11f shows the Freundlich plots for the adsorption of BB on PpPD–Fe3O4. The plot was not linear and this showed that the adsorption process did not follow the Freundlich adsorption model.53
The selectivity of PpPD–Fe3O4 in the photocatalytic degradation of acid dyes can be ascribed to the selective adsorption ability to acid dyes of the catalyst (Fig. 12). The interaction between the acidic groups of dyes and alkalinity groups of PpPD lead to the selective adsorption ability of PpPD–Fe3O4. The experimental results were in agreement with those calculated theoretically. Furthermore, the holes were the main active species in the photocatalytic progress. The enhancement of photoactivity could be attributed to the synergistic effect between Fe3O4 and PpPD. Fe3O4 which have no catalytic activity served as an electron sink and prevented the recombination of photoinduced electrons and holes. The PpPD is excited by UV or visible light and obtained photogenerated carriers and excited electrons, respectively. The excited state electrons in PpPD can readily migrate to the conduction band of Fe3O4. As PpPD is a good material for transporting holes,31,54,55 the photogenerated charges can migrate easily to the surface of the photocatalysts and photodegrade the adsorbed dye molecules.
![]() | ||
Fig. 12 Schematic diagram illustrating the mechanism of adsorption and photodegradation over PpPD–Fe3O4. |
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c4ra11138a |
‡ These authors are co-first authors. |
This journal is © The Royal Society of Chemistry 2014 |