Monika
Wysocka-Żołopa
*a,
Agata
Blacha-Grzechnik
*bc,
Joanna
Breczko
a,
Diana M.
Bobrowska
a,
Adam
Mizera
d and
Krzysztof
Winkler
a
aDepartment of Chemistry, University of Bialystok, Ciolkowskiego 1K, 15-245, Bialystok, Poland. E-mail: monia@uwb.edu.pl
bFaculty of Chemistry, Silesian University of Technology, Strzody 9, 44-100, Gliwice, Poland. E-mail: Agata.Blacha-Grzechnik@polsl.pl
cCentre for Organic and Nanohybrid Electronics, Silesian University of Technology, Konarskiego 22B, 44-100, Gliwice, Poland
dInstitute of Molecular Physics, Polish Academy of Sciences, M. Smoluchowskiego 17, 60-179, Poznań, Poland
First published on 18th August 2025
The polymeric material of PCBM–Pd was prepared using both electrochemical and chemical synthesis. The electrochemical synthesis was carried out in acetonitrile/toluene (1:
4, v/v) solution containing PCBM, palladium acetate, and tetra(n-C4H9)4NClO4 as the supporting electrolyte. The polymeric film was grown at the Au electrode surface under cyclic voltammetry conditions. Chemical synthesis was performed in a benzene solution containing PCBM and Pd2(dba)3·CHCl3. This study investigates singlet oxygen's morphology, structural properties, and photochemical generation in this synthesized n-polymer. The morphology of the PCBM–Pd polymer was observed by scanning electron microscopy (SEM) and transmission electron microscopy (TEM). The diffuse reflectance infrared Fourier transform (DRIFT) spectroscopy, X-ray photoelectron spectroscopy (XPS), thermogravimetric analysis (TGA), and nitrogen adsorption/desorption isotherms (BETs) were used for the structural characterization of this polymeric material. The formation of 1O2 by the photoactive PCBM–Pd polymer was confirmed using 2,3,4,5-tetraphenylcyclopentadienone (TPCPD) as a specific singlet oxygen quencher. Under these conditions, the photogeneration of singlet oxygen differs depending on the molar ratio of PCBM to Pd. The photochemical generation of singlet oxygen for the PCBM–Pd polymer electrochemically deposited on an ITO electrode and gold foil does not depend on the thickness of the deposited layer.
Singlet oxygen, 1O2, is considered an important reactive oxygen species.15 Due to the unoccupied π* orbital presence, 1O2 shows high selectivity towards electron-rich substances and unsaturated organic molecules.16 However, generating triplet states in organic photovoltaic (OPV) cells can be problematic because it reduces the device's efficiency in several ways. First, the presence of triplets lowers the short-circuit current density and open-circuit voltage, resulting in lower overall power conversion efficiency. Triplet excited states can also react with oxygen to form singlet oxygen that easily damages the cells’ photoactive layers, leading to the degradation of the cell and a further reduction in efficiency.17 Thus, the studies on 1O2 generation by photoactive layers applicable in OPVs can provide valuable information on their durability and degradation mechanism. Therefore, analysis of this process allows the design of more stable OPV cells. On the other hand, the high reactivity of singlet oxygen may be helpful in synthesizing specialty chemicals, in wastewater treatment, and in photodynamic therapy (PDT), where its cell-destroying ability is targeted to fight cancer cells.18–20 However, forming singlet oxygen requires certain conditions. Direct excitation of oxygen from the triplet ground state to the singlet excited state is spin-forbidden, meaning that using light directly is impossible. However, it can be achieved using photosensitizers. A photosensitizer first absorbs light, which puts it into a singlet excited state (S1). It then transitions to a triplet excited state (T1) through intersystem crossing (ISC). In the triplet excited state, the photosensitizer can transfer its energy to oxygen in the ground (triplet) state, resulting in the formation of 1O2 and the return of the photosensitizer to the ground state.
Several methods have been developed to detect singlet oxygen, which can be broadly divided into indirect (trapping) and direct spectroscopic techniques. Indirect methods commonly rely on chemical traps, such as 1,3-diphenylisobenzofuran (DPBF) or 2,3,4,5-tetraphenylcyclopentadienone (TPCPD), which react selectively with 1O2 to form stable products that can be monitored by UV-vis absorption spectroscopy.21 These methods are simple and widely used; however, they are prone to interference from side reactions, trap photodegradation, and cannot provide information on the singlet oxygen lifetime or its precise quantum yield.22 Direct detection techniques, in contrast, exploit the characteristic phosphorescence emission of 1O2 at around 1270 nm during the transition from its excited singlet state (1Δg) to the ground triplet state (3Σg−).23 Time-resolved near-infrared (NIR) phosphorescence detection is the gold standard for reliable singlet oxygen formation and decay dynamics measurement.24 This method allows for precise determination of the quantum yield (ΦΔ) and provides insights into the kinetics of 1O2 generation and quenching in different environments.25 In this energy transfer, both singlet excited states of oxygen, 1Σ+g and 1Δg, as well as the triplet ground state 3Σ−g can be formed with different rate constants (k1ΣT, k1ΔT, and k3ΣT). The formation proceeds via two main quenching pathways: one involving non-charge-transfer (nCT) complexes, and another involving partial charge-transfer (pCT) exciplexes. The efficiency of singlet oxygen formation depends on the excess energy in the nCT channel and on the free energy of electron transfer in the pCT channel. Notably, a rapid intersystem crossing equilibrium between singlet and triplet exciplexes appears only in the nCT pathway.26
Additionally, it is known that quenching of ππ excited triplet states by O2 occurs through internal conversion within encounter complexes and exciplexes, leading to singlet oxygen with varying efficiencies. The balance between these deactivation paths depends on the sensitizer's triplet energy, oxidation potential, and solvent polarity.27
The photochemical properties of fullerenes and their derivatives have received intensive attention.28–33 The triplet state of C60 sensitizes singlet oxygen formation in high yields.34 It was shown that electron-deficient fullerenes are useful sensitizers in preparative photooxygenation reactions by the intermediacy of singlet oxygen.35 Unfortunately, the unfunctionalized C60 and C70 molecules possess poor solubility in most common solvents. The functionalization of the fullerene core with solubilizing groups makes the indirect study of their properties in polar solvents possible. For example, a palladium metallosquare with encapsulated fullerenes C60 and C70, which exhibits high solubility in organic solvents such as CH3CN, CH2Cl2, CHCl3, etc., can be used as a photosensitizer for the generation of singlet oxygen and for facilitating carbon–oxygen bond forming reactions with a variety of organic substrates.36 Okutan et al.37 investigated perylenebisimide-fullerene (PBI-fullerene) dyads. These PBI-fullerene dyads were utilized to produce efficient singlet oxygen and were applied in the photooxidation of 1,5-dihydroxy naphthalene (DHN) to produce juglone.37 Photoactive layers based on the P3HT matrix containing carbon nanostructures such as C60, PCBM, and SWCNTs were also studied as photosensitizers for singlet oxygen production.32 The formation of single oxygen was monitored spectroscopically in situ by oxidation of DPBF in methanol and used synthetically to produce juglone. The production yield of 1O2 depended on the carbon nanostructure photosensitizer and was significantly lower for mixtures of P3HT with C60 and single-walled carbon nanotubes than for mixtures of P3HT with PCBM.32
Fullerene moieties can be also bounded in η2-fasion through transition metal-containing linkers to form large macromolecular structures.38 These coordination fullerene polymers can be formed under chemical or electrochemical conditions.38–40 Chemical synthesis takes place in a non-polar solution containing fullerene and a transition metal complex in a low oxidation state.41–44 The polymer precipitates from the solution in the form of large particles composed of spherical nanoparticles with sizes of 50 to 200 nm. Electrochemical synthesis is usually performed under cyclic voltammetry conditions in a solution containing fullerene and transition metal complex. The electrochemical reduction process taking place under such conditions leads to the formation of thin and homogeneous polymer layers on conductive surfaces.45 Coordination polymers in which fullerene moieties were connected by Pd0,45 Pt0,46 IrI (CO)2,45,47 and RhI (CF3COO)245,47 were electrochemically synthesized. Both, pristine fullerenes and fullerene derivatives including porphyrin,48 crown ether,49 ferrocene50,51 and pyrrolidine and piperazine52 derivatives were incorporated into a three-dimensional polymeric structure.
This work reports that the coordination polymer of PCBM and palladium atoms (PCBM–Pd) is synthesized using both chemical42,53,54 and electrochemical45,55,56 C60–Pd formation processes. Such a polymeric material was applied as a photocatalyst to generate singlet oxygen. The PCBM–Pd polymer is practically insoluble in organic and inorganic solvents. Therefore, thin PCBM–Pd layers can be used for singlet oxygen generation in different liquid media. The photoactive performance of such thin films is presented in this work.
The results of electron microscopy studies indicating the dependence of porosity of the formed materials on the conditions of their preparation were confirmed by the results of porosimetric tests. In panel a of Fig. 2, the nitrogen adsorption–desorption curves of chemically synthesized PCBM–Pd polymers with different component mass ratios are shown. They all represent type IV isotherms with hysteresis loops, indicating the dominance of mesopores in the synthesized materials.58,59 The BET (Brunauer–Emmett–Teller) specific surface area (panel b of Fig. 2) of the analyzed polymers increased with increasing palladium content, reaching 22, 92, and 303 m2 g−1 for PCBM–Pd (1:
1), PCBM–Pd (1
:
2), and PCBM–Pd (1
:
3), respectively (Table 1).
Material | BET surface area (m2 g−1) | S micro (m2 g−1) | S external (m2 g−1) | Cumulative volume of poresb (cm3 g−1) | Average pore diameterc (nm) |
---|---|---|---|---|---|
a Calculated using the t-plot method. b BJH adsorption cumulative volume of pores in the diameter range of 17.000–3000.000 Å. c Average pore size diameter calculated using the BJH method. | |||||
PCBM–Pd (1![]() ![]() |
22 | 4 | 18 | 0.0627 | 14.1 |
PCBM–Pd (1![]() ![]() |
92 | 27 | 65 | 0.1327 | 9.3 |
PCBM–Pd (1![]() ![]() |
303 | 56 | 247 | 0.3066 | 7.8 |
The t-plot analysis also indicated that the amount of both micropores and mesopores increased with the growth of palladium content in the obtained polymers (Table 1). Barrett–Joyner–Halenda (BJH) (Table 1) and density functional theory (DFT) (Fig. 6c) calculations confirm a decrease in average pore size with an increase in total pore volume as the mass proportion of PCBM in the polymer material decreases.
IR spectroscopy is widely used to obtain structural information in composition studies and the interaction between components. Fig. 3 shows the spectra of the PCBM–Pd (1:
3) polymer. For comparison, the spectra obtained for pure PCBM, pure C60 fullerene, and C60Pd polymer are also presented.
Pure C60 fullerene gives four characteristic signals at 527, 576, 1180, and 1428 cm−1 (Fig. 3a). The bands at 527 and 576 cm−1 correspond to the deformation vibrations of the C60 molecule, while the bands at 1180 and 1428 cm−1 are attributed to the vibrations of the CC bond in fullerene. In the case of the C60–Pd3 polymer (Fig. 3b), the signals at 486, 526, and 556 cm−1 are attributed to the stretching vibrations of the C–Pd bond, and the signals at 668, 697, and 760 cm−1 are attributed to the deformation vibrations of the C–Pd bond. The peaks at 1183, 1337, 1373, and 1450 cm−1 are characteristic of the vibrations of the C
C region in the benzene ring.39 For the fullerene derivative PCBM (Fig. 3c) at 1630 and 1737 cm−1, signals from stretching and deformation vibrations of the C
O and at 1148 cm−1 from the C–O–C bond in the ester group are also observed.60 On the spectrum obtained for the PCBM–Pd (1
:
3) polymer (Fig. 3d), observed bands from both PCBM and those related to C–Pd bond vibrations indicate this polymer's formation.
In addition, calculations of the oscillatory structure for C60Pd and PBCM–Pd were performed using Gaussian 09 software. The calculations were performed using DFT at the cam-B3LYP/cep-121g theory level.61–64 Vibrational structure calculations were performed for the optimized system. No negative frequencies were observed. Theoretical infrared (IR) spectra were obtained using GaussSum65 software. These spectra were plotted with a half-width of 10 cm and a scaling factor of 0.95. Fig. S1 in the supplementary materials presents theoretical spectra for these systems. Furthermore, theoretical calculations suggest that the bands at 486 and 526 (for C60Pd) and 485 and 524 (for PCBM–Pd) are related to vibrations resulting from the deformation of C–C bonds in the fullerene molecule, which are disturbed by palladium.
X-ray photoelectron spectroscopy (XPS) analysis gave more information about the chemical composition of the chemically prepared materials. In Fig. 4, the XPS spectra of pure PCBM are presented. The high-resolution XPS spectrum of the C 1s region can be deconvoluted into several peaks corresponding to carbon atoms in different functional groups present in the analyzed sample (Fig. 4a). This spectrum shows a prominent peak at 284.8 eV, which originates from the C–C/CC bond in C60 and the phenyl group. The other components located at 286.4 eV and 288.7 eV are characteristic of epoxy (C–O) and carbonyl (C
O/O–C
O) groups, respectively, and arise from the butyric acid methyl ester side chain of PCBM. The broad peak at 290.8 eV can be assigned to a shake-up feature involving the π → π* transition in the C60.66,67 The decomposition of the O 1s spectrum gives two components about the methyl ester group, as shown in Fig. 4b. The peak at 531.9 eV corresponds to carbonyl groups (C
O/O–C
O). The peak at 533.9 eV corresponds to the epoxy group (C–O).67,68
Fig. 5 shows the C 1s, O 1s, and Pd 3d XPS high-resolution spectra obtained for PCBM–Pd polymers. The sample C 1s spectrum and O 1s spectrum recorded for PCBM–Pd (1:
1) are shown in Fig. 5a and b, respectively. The C 1s spectrum recorded for PCBM–Pd obtained for different molar ratios exhibits the same components as pure PCBM (Fig. 4a). In the case of O 1s, compared to PCBM, an additional signal at 531.0 eV assigned to Pd–O can be observed. The analysis of the high-resolution XPS spectrum of the Pd 3d region reveals two doublet signals (Fig. 5c–e). The first one, located at 336.6 and 341.9 eV, represents Pd 3d5/2 and Pd 3d3/2, respectively, and it can be assigned to palladium atoms bonded to fullerene moieties.69,70 The second pair of peaks at 338.3 and 343.75 eV also appear in the Pd 3d spectrum, which agrees with the values reported for Pd(IV).71–73 Studies using TEM confirmed the presence of Pd nanoparticles on the surface of the synthesized PCBM–Pd polymer, with their amount increasing with increasing concentration of the precursor during the polymerization process. Thus, there is a possibility of oxidation of metallic Pd nanoparticles to PdO2 during the sample drying process.72 Hence, signals originating from Pd(IV) may appear.
The TGA stability curves of PCBM and polymers of PCBM–Pd in an air atmosphere are presented in Fig. 6. PCBM shows stability up to about 400 °C. In a temperature range from 400 to 600 °C, PCMB undergoes very rapid degradation associated with the decomposition of C60 molecules. When 600 °C is reached, PCBM is almost completely volatilized, leaving a total residue of ca. 0.3%. This result is consistent with the literature on TGA thermograms for PCBM in an air atmosphere.74 PCBM–Pd polymers begin to degrade slightly earlier at about 300 °C. Still, at 700 °C, the value of total residual palladium mass is equal to 10.53, 22.68, and 30.41 wt% for PCBM–Pd (1:
1), PCBM–Pd (1
:
2), and PCBM–Pd (1
:
3), respectively. The PCBM to Pd molar ratios determined from the TGA curves are equal to 1
:
0.95, 1
:
2.08, and 1
:
2.85, for materials formed in solutions with a PCBM to Pd ratio of 1
:
1, 1
:
2, and 1
:
3, respectively. A good agreement of the PCMB to Pd molar ratio in the grown solution and the formed material indicates 1-D, 2-D, and 3-D polymeric structures depending on the solution composition, as was observed for the process of C60–Pd formation.42 However, in the case of PCMB, one of the sites available in the carbon lattice for binding palladium is blocked by an ester moiety. A theoretical value of the PCMB to Pd molar ratio of 1
:
2.5 should be expected in this case. Slightly higher values obtained from thermogravimetric measurements indicate the formation of palladium nanoparticles, as indicated by the results of XPS studies discussed earlier.
Therefore, the dominant degradation pathway of PCBM–Pd should involve singlet oxygen, which is generated from triplet excited states located at PCBM.78,79 For PCBM, triplets can be formed via intersystem crossing from the PCBM singlet states after light excitation, before being quenched by oxygen via energy transfer, forming singlet oxygen.79
Fig. 7 shows UV-vis absorption spectra of the PCBM–Pd (1:
3) polymer and TPCPD in 1,2-dichlorobenzene solution. The UV-vis spectrum of pristine PCBM in 1,2-dichlorobenzene is also shown for comparison. The spectrum in the UV range below 350 nm, both for pure PCBM and polymer of PCBM–Pd, corresponds to the strong allowed transition of the fullerene core.80 In the visible region, where the transition is forbidden mainly by the symmetry of the highly symmetric C60 molecule, much more prominent peaks are observed in the range from 400 to 700 nm in pure PCBM and PCBM–Pd, suggesting that the fullerene derivative side chain only interferes with the forbidden transitions of the fullerene core in the visible region as a result of breaking the symmetry of the original C60 molecule by assigning a [6,6]-addition to C60.80,81 The UV-vis spectra of the TPCPD solution show a characteristic absorption peak in the range of 420–700 nm. Absorption is assumed to decrease TPCPD at 510 nm and is directly proportional to its reaction with singlet oxygen.
Fig. 8 presents the UV-vis spectra of a 0.01 mM solution of TPCPD in 1,2-dichlorobenzene without photosensitizer of PCBM or PCBM–Pd recorded every 2 minutes during the irradiation of this solution for 20 minutes. As you can see, small changes in the absorption band intensity at 510 nm indicate that this molecule is stable under light illumination with a xenon lamp, can be used in 1,2-dichlorobenzene, and is selective toward 1O2. Therefore, the pure solution of TPCPD excludes its self-decomposition under lighting.
![]() | ||
Fig. 8 UV-vis absorption spectra of a 0.01 mM solution of TPCPD in 1,2-dichlorobenzene upon illumination with a xenon lamp as a light source for 20 minutes. |
Fullerene electron acceptors are widely used in organic electronics, and PCBM is the most commonly employed electron acceptor in organic solar cells. This is down to the low-lying excited states of the fullerene anion, which promote fast charge separation and high solubility in organic solvents, making solvent processing possible.82 Most studies use PCBM as an acceptor when testing new polymers, architectures, or upscaling methods. However, despite being well established, the photo-oxidation of fullerene electron acceptors is far less studied than that of polymers. The photochemical generation of singlet oxygen, 1O2, using PCBM typically involves sensitizing molecular oxygen (O2) by a photosensitizer, such as PCBM, under light irradiation. PCBM absorbs light, typically in the visible or near-infrared region. The absorption of a photon promotes the sensitizer to an excited state. PCBM can undergo intersystem crossing to a triplet state in the excited state. The triplet state of PCBM can then transfer its energy to ground-state molecular oxygen, O2, producing singlet oxygen. The energy transfer process involves the transfer of energy from the excited state of PCBM to ground-state molecular oxygen, resulting in the formation of singlet oxygen. PCBM sensitizes the molecular oxygen in its ground state (3O2) to produce 1O2. The overall reaction can be represented as:
PCBM + hν → PCBM* | (1) |
PCBM* + O2→ PCBM + 1O2 | (2) |
Fig. 9a–c shows the sets of UV-vis spectra of TPCPD in 1,2-dichlorobenzene recorded during irradiation of the chemically synthesized PCBM–Pd (1:
3) polymer added in varying amounts: (a) 50 μl, (b) 100 μl, and (c) 150 μl. In all materials tested, a decrease in the absorption band located at 510 nm was observed with increased irradiation time of the PCBM–Pd polymer. These drops in the absorption band indicate the reaction of TPCPD with singlet oxygen photogenerated by the polymer of PCBM–Pd. The comparison of the decrease in the intensity of the TPCPD band at 510 nm over time (Fig. 9d) confirms a much faster oxidation of TPCPD with an increase in the polymer's concentration in solution. UV-vis absorption spectra obtained for pure PCBM as a photosensitizer dissolved in various amounts in a solution of 1,2-dichlorobenzene containing 0.01 mM TPCPD were very similar to those obtained in the case of the PCBM–Pd polymer synthesized in a PCBM to Pd2(dba)3·CHCl3 molar ratio of 1
:
1 (Fig. S2 in SI). An increase in the photooxidation rate of the polymer is also observed as its concentration in the solution increases. Thus, the studies indicate that the PCBM in the polymer is the source of singlet oxygen. An increase in the ratio of Pd concentration to PCBM concentration during the chemical synthesis of the PCBM–Pd polymer leads to much slower changes in absorbance associated with the photooxidation of this polymer. This is due to the decrease in the amount of the fullerene derivative in the polymer.
Reaction rate constants for different amounts of PCBM–Pd polymers in solution as a source of active oxygen were determined by linear regression as the slope of the line and are presented in Table 2. These changes related to the different rates of decrease in the absorption bands for polymers with different PCBM to Pd2(dba)3·CHCl3 concentration ratios depend on the amount of PCBM in the sample. For the PCBM–Pd system (1:
1), where there are more PCBM in one volume, the changes in these ratios are much higher compared to PCBM–Pd (1
:
3). In addition, the rate of singlet oxygen generation may also be related to the structure of the polymer being formed. In the case of agglomerated spherical particles, which form as the ratio of palladium to PCBM increases, it is much more difficult for oxygen to penetrate more aggregated areas. The photooxidation of fullerene derivative polymers can, therefore, be altered by changing their concentration and structure by increasing their crystallinity or aggregation, and can be used to control these properties.
Photoactive polymer | k [mM min−1] |
---|---|
50 μl PCBM–Pd (1![]() ![]() |
6.92 × 10−2 |
100 μl PCBM–Pd (1![]() ![]() |
9.72 × 10−2 |
150 μl PCBM–Pd (1![]() ![]() |
15.86 × 10−2 |
50 μl PCBM–Pd (1![]() ![]() |
4.19 × 10−2 |
100 μl PCBM–Pd (1![]() ![]() |
6.50 × 10−2 |
150 μl PCBM–Pd (1![]() ![]() |
9.03 × 10−2 |
50 μl PCBM–Pd (1![]() ![]() |
2.86 × 10−2 |
100 μl PCBM–Pd (1![]() ![]() |
4.09 × 10−2 |
150 μl PCBM–Pd (1![]() ![]() |
6.60 × 10−2 |
A similar set of UV-vis spectra was obtained during irradiation of the PCBM–Pd layer of different thicknesses, electropolymerized onto the ITO electrode and thin gold foil (Fig. S3 in SI). Examples of UV-vis spectra of PCBM–Pd electrochemically synthesized on gold foil are shown in Fig. 10. The studies carried out show that the number of cycles, and thus the amount of deposited polymeric material, does not affect the rate of the photooxidation layers, because it is a typical surface process. However, much better adhesion can be seen in the case of layers deposited on gold foil.
The morphology of the obtained materials was secondary-electron imaged using an Inspect S50 scanning-electron microscope (FEI Company, Hillsboro, Oregon, USA). The accelerating voltage of the electron beam was 20 or 25 keV, and the average working distance was 10 mm.
TEM images were obtained using a Tecnai G2 20 X-TWIN microscope (FEI Company, Hillsboro, Oregon, USA) with a LaB6 emitter and an HAADF detector operating at 120/200 kV. Ametek's EDX instrument and software (Berwyn, Pennsylvania, USA) were used for energy-dispersive X-ray fluorescence measurements. This software facilitates standardless quantification of atomic weight and atomic percentage. The accelerating voltage for the electron beam was 20 keV, and the working distance was 10 mm.
The DRIFT spectra were recorded using a Magna IR 550 Series II spectrometer with a spectral resolution of 4 cm−1.
Thermogravimetric analysis (TGA) was performed using a Mettler Toledo Star TGA/DSC system. Nitrogen was used as the purge gas (0.1 dm3 min−1). Samples weighing 2 mg were placed in aluminum pans and heated from 50 to 1000 °C at a heating rate of 10 °C min−1.
Nitrogen adsorption/desorption isotherms (BETs) were recorded at 77.15 K with a Micrometrics ASAP 2020 automatic sorption analyzer.
X-ray photoelectron spectroscopy (XPS) analysis was conducted using an AXIS Supra+ instrument (Kratos Analytical) equipped with a monochromatic Al Kα X-ray source (hν = 1486.6 eV, operating at 10 mA, 15 kV). The system base pressure was equal to pb = 1.7 × 10−9 Torr. The pass energy was set to 160 eV (scanning step 0.9 eV) or 20 eV (scanning step 0.05 eV) for survey spectra and high-resolution spectra, respectively. For the charging effect compensation, the Kratos charge neutralizer system was used. The binding energy scale was calibrated concerning the C–C component of C 1s spectra (284.8 eV). The acquired spectra were analysed using CASA XPS® software and embedded algorithms. The components of the high-resolution spectra were presented with Gaussian (70%) and Lorentzian (30%) lines, while the background was with Shirley's function.
Singlet oxygen (1O2) photogeneration by PCBM–Pd3 was investigated using a Hewlett–Packard 8452A UV-vis spectrometer and a standard 10 mm × 4 mm quartz cuvette (Hellma Analytics). A xenon lamp was employed as an excitation light source and was arranged perpendicular to the direction of UV-vis spectra acquisition.
Electrochemical synthesis was performed in acetonitrile/toluene (1:
4, v/v) solution containing 0.22 mM PCBM, 3.56 mM palladium acetate, and 0.1 M tetra(n-C4H9)4NClO4 as the supporting electrolyte on the ITO electrode and gold foil electrode. The film was grown at the electrode surface under cyclic voltammetry conditions in the potential range of −0.2 to −1.05 V at a potential sweep rate of 100 mV s−1 (Fig. S4 in SI). The increase in current values in subsequent cycles indicates the formation of a new solid phase on the electrode surface. Therefore, in this case, the polymer was deposited at the electrode surface during electropolymerization. The thickness of the deposited polymer layer depended on the number of cycles used in cyclic voltammetry.
Chemical synthesis was performed in benzene solution containing 0.48 mM PCBM and respectively 0.24 mM, 0.48 mM, or 0.73 mM Pd2(dba)3·CHCl3 according to the procedure proposed by Nagashima,42 obtaining a polymer with the appropriate molar ratio of PCBM to Pd2(dba)3·CHCl3 of 1:
1, 1
:
2, 1
:
3 (PCBM–Pd (1
:
1), PCBM–Pd (1
:
2), PCBM–Pd (1
:
3)), respectively. By chemical synthesis, a black polymeric material was obtained.
The photocatalytic activity of PCBM–Pd toward singlet oxygen evolution was examined. Photogeneration of singlet oxygen in phenyl-C61-butyric acid methyl ester (PCBM) can occur through the photosensitization process in the presence of TPCPD as a specific 1O2 quencher. This work also represents the first demonstration of singlet oxygen generation and the specific photochemical sensitizing ability of the PCBM–Pd polymer. This photoactive polymer can, therefore, be considered as an alternative in the heterogeneous oxidation reactions activated by visible light.
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2025 |