Next Article in Journal
Molybdenum Carbide/Ni Nanoparticles Embedded into Carbon Nanofibers as an Effective Non-Precious Catalyst for Green Hydrogen Production from Methanol Electrooxidation
Previous Article in Journal
Anticorrosion, Thermal Degradation, and Hydrophobic Performances of Graphene/TiO2 Nanocomposite Coatings
Previous Article in Special Issue
Synthesis of a Dual-Color Fluorescent Dendrimer for Diagnosis of Cancer Metastasis in Lymph Nodes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Characterization of the Donor-Acceptor Conjugated Polymer PBDB-T Implementing Group IV Element Germanium

Department of Chemistry, Delaware State University, Dover, DE 19901, USA
*
Author to whom correspondence should be addressed.
Polymers 2023, 15(11), 2429; https://doi.org/10.3390/polym15112429
Submission received: 20 March 2023 / Revised: 20 May 2023 / Accepted: 22 May 2023 / Published: 23 May 2023

Abstract

:
Here, we synthesized and characterized a novel two-dimensional (2D) conjugated electron donor–acceptor (D-A) copolymer (PBDB-T-Ge), wherein the substituent of triethyl germanium was added to the electron donor unit of the polymer. The Turbo–Grignard reaction was used to implement the group IV element into the polymer, resulting in a yield of 86%. This corresponding polymer, PBDB-T-Ge, exhibited a down-shift in the highest occupied molecular orbital (HOMO) level to −5.45 eV while the lowest unoccupied molecular orbital (LUMO) level was −3.64 eV. The peaks in UV-Vis absorption and the PL emission of PBDB-T-Ge were observed at 484 nm and 615 nm, respectively.

Graphical Abstract

1. Introduction

Polymer solar cells (PSCs), comprising bulk-heterojunction (BHJ) photoactive layers of electron donors (D) and electron acceptors (A), have attracted enormous attention due to their lightweight properties, solution-processability, low-temperature printing fabrication and mechanical flexibility [1,2,3,4,5,6,7,8]. Several types of electron-acceptor materials have been developed for use in BHJ PSCs, including fullerene and non-fullerene derivatives, which cover allotropes, polymers, and small organic molecules (SMAs) [7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24]. Fullerene derivatives were reported to exhibit attractive properties that made them promising electron acceptors for PSCs in the past. These characteristics of the fullerene derivatives include affordable three-dimensional electron mobility, forming appropriate phase-separation in the photoactive layer [24,25,26]. However, despite the outstanding performance of fullerene-based PSCs, fullerene derivatives still exhibit many intrinsic drawbacks such as high production costs with tedious purification steps, weak absorption in the visible spectral region, and the limited tunability of energy level [24,25,26].
Other electron acceptors, such as SMAs, have shown advantages in simple synthesis methods with various structural variations to acquire favorable absorption ranges and modulate frontier energy levels [24,25,26,27,28,29]. Due to the advantages of SMAs, the power-conversion efficiencies (PCEs) of the SMA-based PSCs were reported to have continuously improved by ca. 17% [6,30,31,32]. Considerable efforts have been made to construct high-performance non-fullerene (NF) PSCs that match the electronic band structure of electron donors and electron acceptors to absorb photons in the visible and near-infra red regions. The structural modification of electron donors and acceptors can improve short-circuit currents (Jsc), and the alignment of the energy levels between electron donors and acceptors was reported to control open-circuit voltage (Voc). The crystallinity and morphology of electron donor–acceptor pairs have also been reported to impact the fill factor (FF) [7,32,33].
PSCs use the variable bandgap conjugated polymer to construct the BHJ photoactive layer, modulating electron-donating (D) and electron-accepting (A) characteristics of the conjugated D-A polymers. The molecular engineering of D-A moieties in the backbone of the polymers is considered to be one of the most successful strategies to tune the energy levels and adjust optical properties [4,33]. Recently, many types of D-A copolymers have been developed for the photoactive layer of PSCs and reported to show a PCE of 16% or higher [1,30].
Recently, Bin H. et al. adopted a new strategy to create copolymers, tuning the highest occupied molecular orbital (HOMO) by implementing the element Si, one of the group IV elements, in the backbone of the BDTT-FBTA (J71) copolymer [34]. Replacing the alkyl group with the trialkyl silyl substituent decreased the HOMO to −5.4 eV, enhancing the absorption, which was reflected by the achievement of a PCE of 11.4%. Previously, analogous alkyl sidegroup-based PSC exhibited a PCE of 9.5%. To examine this idea, Chen S. et al. synthesized several D-A copolymers, coupling BDTT-Si with different acceptor moieties. The PSCs were reported to show a PCE of 12.1% [33].
Motivated by previous studies that implemented the group IV elements in the backbone of the D-A-conjugated copolymers, we investigated another group IV element, germanium. It is expected to have stronger π-π stacking due to the longer C–Ge bond length compared with the C–C and C-Si bonds. It is known to assist in intermolecular charge transport and suppress the charge carrier recombination, which eventually improves the photovoltaic performance. Here, we have synthesized and characterized a new copolymer, PBDB-T-Ge, using the substituent of trialkyl germane on the BDTT (BDTT-Ge) moiety, which is the donor unit of the copolymer and is coupled with the acceptor moiety BDD. Figure 1 illustrates the molecular structures of PBDB-T-Ge and precedent copolymers, PBDB-T-C and PBDB-T-Si, where the sidegroups of the electron donor of the polymers are trialkyl germane, ethyl hexyl and tripropyl silyl substituent, respectively. In this paper, we focus on the synthesis and characterization of the PBDB-T-Ge to compare the electronic band structure.

2. Materials and Experimental Methods

2.1. Materials

Dichloromethane (DCM, 99.9%), isopropyl magnesium chloride lithium chloride complex (iPrMgCl·LiCl, 1.3 M in THF), trimethyl tin chloride (1 M in hexane, 25 wt.%), acetone and deuterated chloroform (dCDCl3) were purchased from Acros Organics (Branchburg, NJ, USA). Hexane, methanol, toluene, hydrochloric acid (HCl), chloroform, acetonitrile and ethyl acetate were purchased from Fisher Scientific (Branchburg, NJ, USA). Ferrocene, tetraethylammonium tetrafluoroborate (Et4NBF4) and diethyl ether were obtained from Alfa Aesar (Ward Hill, MA, USA). Anhydrous tetrahydrofuran (THF), n-Butyllithium solution (n-BuLi, 2.5 M in hexane), Benzo [1,2-b:4,5-b′]dithiophene-4,8-dione, tin(ll) chloride dihydrate, anhydrous sodium sulfate (>99%), indium tin oxide (ITO) glass substrate, 1,3-Bis(5-bromo-2-thienyl)-5,7-bis(2-ethylhexyl)-4H,8H-benzo [1,2-c:4,5-c′]dithiophene-4,8-dione (BDD, 97%), tetrakis(triphenylphosphine) palladium (Pd(PPh3)4) and Y6 were purchased from Sigma-Aldrich (Burlington, NJ, USA). Triethyl germanium chloride and 2-iodothiophene were obtained from TCI (Boston, MA, USA). Ethanol was obtained from EMD (Burlington, NJ, USA). Isopropyl alcohol was purchased from VWR (Missouri City, TX, USA). All materials were used as received without further purification.

2.2. Synthesis of PBDB-T-Ge

2.2.1. Synthesis of Triethyl (Thiophen-2-yl) Germane (2)

Triethyl(thiophen-2-yl) germane was synthesized via a method that was adopted from Sherborne G. et al. [35]. Under the protection of argon, THF (141 mL) was placed in a 500 mL, two-neck, round-bottom flask. Triethyl germanium chloride (5.5 g, 4.7 mL, 28.16 mmol) and 2-iodothiophene (5.38 g, 2.83 mL, 25.6 mmol) (1) were added to the round-bottom flask under stirring condition. Then, iPrMgCl·LiCl (23.6 mL, 30.73 mmol, 1.3 M in THF) was added dropwise to the reaction mixture, which was left under stirring conditions for 3 h at room temperature. The mixture was quenched with cold DI water (30 mL), which was then extracted by DCM (25 mL) three times. The organic layer was collected and dried over anhydrous sodium sulfate (Na2SO4). The volatile solvent was removed under vacuum with mild temperature; then, the residue was purified by column chromatography using hexane: ethyl acetate (5:1 v/v) as an eluent with a retention factor (RF) of 0.0789 to yield the triethyl(thiophen-2-yl) germane as a colorless oil. Yield = 5.4 g (86%). 1H NMR (400 MHz, CDCl3), δ (ppm): 7.62–7.59(d, 1H), 7.25–7.22(m, 1H), 7.20–7.19(d, 1H), 1–1.3(m, 15H).

2.2.2. Synthesis of 4,8-Bis(5-triethyl germanium)thiophene-2-yl)benzo[1,2-b:4,5-b′]dithiophene (3)

Under the protection of argon, a solution of triethyl(thiophen-2-yl) germane (2.5 g, 10.29 mmol, compound 2) in THF (45 mL) was placed in a 100 mL, two-neck, round-bottom flask. The solution was cooled to 0 °C. n-BuLi (4.12 mL, 10.29 mmol, 2.5 M in hexane) was added dropwise into the RBF and the mixture was kept at 0 °C for 40 min before the mixture was refluxed at 65 °C under stirring for 2 h. After we cooled the mixture down, 4,8-dehydrobenzo[l,2-b:4,5-b′] dithiophene-4,8-dione (0.75 g, 3.43 mmol) was added in one portion, and the mixture was refluxed with stirring for another 2 h. Tin(ll) chloride dihydrate (6.19 g, 27.44 mmol) in 10% HCl (15 mL) was added after the mixture was cooled down to room temperature and stirred overnight. The reaction was quenched using cold DI water (40 mL), and then the crude product was extracted using diethyl ether (25 mL) three times. The organic layer was collected and dried over anhydrous sodium sulfate (Na2SO4). The volatile solvent was removed under vacuum at a mild temperature and the crude product was purified using a column chromatography with hexane: ethyl acetate (5:1 v/v) as eluents to yield 4,8-bis-(5-(triethyl germanium)thiophene-2-yl)benzo[1,2-b:4,5-b′]dithiophene as bright yellow gel. Yield = 0.8 g (36.2%). 1H NMR (400 MHz, CDCl3), δ (ppm): 7.67–7.65(d, 2H), 7.6–7.58(d, 2H), 7.47–7.45(d, 2H), 7.3–7.28(d, 2H), 1.00–1.4(m, 30H).

2.2.3. Synthesis of [4,8-Bis[5-triethyl germanium) thiophene-2-yl]-2-trimethylstannyl thieno[2,3-f][1]benzothiol-6-yl]-trimethylstannane (4)

Under the protection of argon, a solution of 4,8-bis-(5-(triethyl germanium) thiophene-2-yl)benzo[1,2-b:4,5-b′]dithiophene (0.4 g, 0.596 mmol) in THF (6 mL) was placed in a 25 mL round-bottom flask. The solution was cooled to −78 °C using dry ice with acetone, and then n-BuLi (0.596 mL, 1.49 mmol, 2.5 M in hexane) was added dropwise. After the addition, the mixture was kept at −78 °C for 1 h; trimethyl tin chloride (1.78 mL, 1.78 mmol, 1 M in hexane) was added dropwise. The resulting mixture was stirred overnight at room temperature. The mixture was then poured into cold DI water (6 mL) and extracted with diethyl ether (8 mL) three times; after drying over anhydrous sodium sulfate (Na2SO4), the volatile solvent was removed under vacuum at a mild temperature. Recrystallization was used for purification using ethanol as a solvent, and then dried for two days in a vacuum at a mild temperature to yield a yellowish white crystal of [4,8-Bis[5-triethyl germanium)thiophene-2-yl]-2-trimethylstannylthieno[2,3-f][1]benzothiol-6-yl]-trimethylstannane. Yield = 0.092 g (15.6%). 1H NMR (400 MHz, CDCl3), δ (ppm): 7.49(s, 2H), 7.48–7.46(d, 2H), 7.2– 7.3(d, 2H), 1.2–0.9(m, 30H), 0.6–0.5(m, 18H).

2.2.4. Synthesis of PBDB-T-Ge (6)

In a 25 mL, double-neck, round-bottom flask, [4,8-Bis[5-triethyl germanium)thiophene-2-yl]-2-trimethylstannylthieno[2,3-f][1]benzothiol-6-yl]-trimethylstannane (50 mg, 0.08 mmol) and 1,3-Bis(5-bromo-2-thienyl)-5,7-bis(2-ethylhexyl)-4H,8H-benzo[1,2-c:4,5-c′]dithiophene-4,8-dione (58 mg, 0.08 mmol) were dissolved in dry toluene (5 mL). The reaction container was purged with argon for 20 min, and then tetrakis(triphenylphosphine)palladium (Pd(PPh3)4) (12 mg, 0.01 mmol) was added. After another flushing with argon for 20 min, the reactants were heated up to reflux them at 120 °C for 24 h. The solution was cooled down to room temperature and poured into cold methanol (MeOH) (50 mL), and then washed three times using cold methanol. The product was dried under vacuum at a mild temperature for 24 h to yield a dark purple solid of PBDB-T-Ge. Yield 0.09 g. Scheme 1 illustrates the synthetic route of PBDB-T-Ge. Mn and Mw of PBDB-T-Ge were measured to be 2.8 KDa and 3.9 KDa using GPC (CHCl3), respectively. The polydispersity (PDI) was observed to be 1.41. 1H NMR (400 MHz, CDCl3), δ (ppm): 7.6–7.75(d, 2H), 7.45–7.35(d, 2H), 7.2–7.3(d, 2H), 7.15–7.25(d, 2H), 6.95–7.05(t, 2H), 1.4–0.95(m, 30H), 0.9–0.6(m, 34H).

2.3. Characterization

The structure of PBDB-T-Ge was confirmed using an NMR spectrometer (Varian Unity Inova, 400 MHz). The spectra were collected using d-chloroform as solvent and the internal chemical shift reference standard was selected based on the residual proton resonance of the solvent, which were acquired, processed, and analyzed using Bruker Topspin (4.1.4) software. The molecular weight of polymers was determined by gel permeation chromatography (GPC) relative to polystyrene standards with chloroform as the eluent at a concentration of 1.0 mg/ mL for at least 12 h. The optical properties were tested using a UV-Vis spectrometer (Shimadzu 3600 Plus). This characterization was performed using chloroform solution in a quartz cuvette using polymer film on a glass substrate. The polymer film samples were prepared using drop-casting on a glass substrate. The photoluminescence study was performed using Spectro-fluorophotometer (Shimadzu RF-5301PC model) using a chloroform solution. Cyclic voltammetry study was performed using a potentiostat (CH instrument electrochemical workstation, model CHI 660C) at room temperature to determine the redox reaction. The surface morphology characterizations on PBDB-T-Ge:Y6 (1:1.2, w/w) composite film were performed using atomic force microscopy (AFM, Bruker Inova) using a taping mode for imaging. Further surface morphology was performed using scanning electron microscopy (SEM, FEI. Quanta FEG 250) with secondary electron for the imaging.

3. Results and Discussion

3.1. Synthesis of PBDB-T-Ge

Our work implemented germanium on the side branch of the PBDB-T-Ge. Previous reports utilized n-butyl lithium to activate the ethylhexyl branch groups of PBDB-T-C and trialkyl silyl groups of PBDB-T-Si, as seen in Figure 1. Although this method is successful for alkylation on the backbone of polymers, it requires an extremely low temperature of −78 °C and an extended reaction time with overnight stirring [34,36]. In this report, we adopted a new method to synthesize PBDB-T-Ge, introducing a sidegroup of trialkyl germanium using a Turbo–Grignard reagent (iPrMgCl·LiCl), which allowed for alkylation to occur at room temperature and the reaction time was cut down to three hours [35]. Additionally, using the Turbo–Grignard reaction led to a higher yield of 86%, while the conventional alkylation methods (BuLi method) produced a lower yield of 45.3%. The Stille C-C polycondensation reaction was used to polymerize PBDB-T-Ge using comonomers of BDD and BDTT-Ge, which are an electron-acceptor and electron-donor, respectively [37].
1H NMR spectra in Figure 2 show that triethyl(thiophen-2-yl) germane (2) is synthesized, where the three peaks between 7.1 and 7.7 ppm are assigned to the proton of the single-branched thiophene and the multiple peaks from 0.8 to 1.4 ppm are assigned to the alkyl groups. The 1H NMR spectra show that the implementation of triethyl germanium on the thiophene was successful, simplifying and enhancing product yields.

3.2. UV-Vis Absorption of PBDB-T-Ge

Figure 3 shows the UV-visible absorption characteristics of comonomers and polymers, which are BDTT-Ge (M1), BDD (M2), PBDB-T-Ge, respectively. In a chloroform solution, the comonomers BDTT-Ge and BDD showed a defined absorption profile with single peaks at 385 nm and 445 nm, respectively, while PBDB-T-Ge exhibited two peaks-absorption profiles at 359 and 484 nm. The red shift in the maximum absorption peaks of the co-monomers confirms successful polymerization, showing extended conjugation.
The UV-Vis absorption spectra of the PBDB-T-Ge solution in chloroform and the polymer film on a glass substrate were compared, as shown in Figure 4, where the absorption of the polymer solution appears to be between 410 and 630 nm with an absorption onset (λonset) at 620 nm, and the maximum absorption (λmax) occurred at 477 nm. In comparison, the absorption peak of reference polymer (PBDB-T-C, Figure 1) with alkyl substituent [38] was reported to be between 500 and 650 nm, showing a maximum absorption peak at 618 nm. The absorption peak of PBDB-T-Ge film was found to appear between 420 and 680 nm, with an onset (λonset) of 666 nm and maximum absorption (λmax) at 507 nm, while another reference polymer film (PBDB-T-Si, in Figure 1) [33] showed an absorption peak between 500 and 670 nm. The UV-Vis absorption peak of the commercial electron acceptor, Y6, in thin film appeared between 580 and 920 nm, with λmax of 838 and λonset of 920 nm. PBDB-T-Ge and Y6 seems to exhibit a complementary absorption, which is beneficial for light harvesting, making them a good candidate for use in a blend film in non-fullerene PSC. The optical properties of PBDB-T-Ge are summarized in Table 1, along with the emission profile. The onset absorption and maximum UV-Vis absorption in solution and the film for PBDB-T-Ge and Y6 are shown in Figure 4. It is noticeable that there is a red shift of 30 nm from solution to film, indicating that there are stronger π-π intermolecular interactions in the film with a better stacking structure.

3.3. Photoluminescence of PBDB-T-Ge

PBDB-T-Ge solution showed a red–orange–fluorescent light at an excitation of 470 nm, as seen in Figure 5. The emission peak appeared in the range between 550 and 740 nm, with the maximum emission peak shown at 615 nm. The intensity of the emission peak in the polymer was observed to be suppressed by the addition of the electron-acceptor PC60BM solution in chloroform, as shown in Figure 5. The fluorescence-quenching explains that the photo-induced charge transfer occurred between the electron-donating PBDB-T-Ge and the electron-accepting PC60BM.

3.4. Electrochemical Characteristics of PBDB-T-Ge

The HOMO and LUMO of PBDB-T-Ge were measured to evaluate the effect of the substitution of the triethyl germane sidegroup on the electronic energy levels using cyclic voltammetry. The electrochemical study was performed using conventional three electrodes configuration consisting of an Ag/AgCl electrode as a reference electrode, the glassy carbon electrode as the working electrode, and the platinum wire as the counter-electrode. We used tetraethylammonium tetrafluoroborate (Et4NBF4, 0.1 M) in acetonitrile as the supporting electrolyte and the ferrocene/ferrocenium as the internal reference. Cyclic voltammogram was obtained at a scan rate of 10 mV/s with a 1 mV sample interval.
Solutions of PBDB-T-Ge polymer and Y6 were prepared by dissolving in chloroform and the films were deposited onto the glassy carbon button electrode via a drop-casting procedure. It was then air-dried. HOMO and LUMO energy levels were estimated from onset oxidation and reduction potentials (Eoxonset and Eredonset), respectively, according to the Equations (1) and (2) [36].
E (HOMO) = −e [Eoxonset + 4.8 − E1/2.Fc/Fc+]
E (LUMO) = −e [Eredonset + 4.8 − E1/2.Fc/Fc+]
where E1/2.Fc/Fc+ is the half-potential of ferrocene/ferrocenium coupled to the electrochemical measurement system, and the energy level of Fc/Fc+ was taken as 4.8 eV below vacuum. Ferrocene/ferrocenium half-potential was measured to be 0.44 V ferrocene/ferrocenium, and Equations (3) and (4) are as follows:
E (HOMO) = −e [Eoxonset + 4.36]
E (LUMO) = −e [Eredonset + 4.36]
Figure 6a shows the cyclic voltammogram of PBDB-T-Ge film, from which the onset oxidation potential (Eoxonset) and onset reduction potential (Eredonset) of PBDB-T-Ge were measured to be 1.09 and −0.72 V vs. Ag/AgCl, respectively. The HOMO energy level and LUMO energy level of PBDB-T-Ge were calculated to be −5.45 and −3.64 eV, respectively, as seen in Figure 6b. Under the same experimental conditions, the HOMO and LUMO energy levels of Y6 were found to be −5.65 eV and −3.86 eV, respectively. The LUMO energy offsets between PBDB-T-Ge as donor and Y6 as acceptor were larger than the exciton-binding energy, which ensures efficient photo-induced electron transfer.
We noticed that the HOMO of PBDB-T-Ge was deepest among the three polymers and showed a smaller gap of 0.22 eV and 0.31 eV compared to PBDB-T-C and PBDB-T-Si, respectively. The small band gap is well-known to improve Jsc, which is expected to enhance the photovoltaic performance. The data were collected from PBDB-T-C in 0.1 M tetrabutylammonium hexafluorophosphate (Bu4PF6) acetonitrile solution at a scan rate of 50 mV/s, while the data of PBDB-T-Si were collected using Ultraviolet Photoelectron Spectroscopy (UPS). The detailed electrochemical data for PBDB-T-Ge and its counterparts, Y6, are listed in Table 2, which indicates that the triethyl germane substitution with σ* (Ge) – π* (C) bond interaction is an effective way to lower the HOMO energy level of 2D-conjugated polymers. Lowering the HOMO energy level is beneficial to improve Voc, which is expected to enhance the PCE of PBDB-T-Ge-based PSC. It should be noted that the electrochemical bandgap (Egcv) of PBDB-T-Ge is 1.81 eV, which is similar to its optical bandgap of 1.86 eV.

3.5. Morphology Study

Molecular-disorder-induced nanostructure variations in the BHJ are critical in optimizing photo-induced charge separation and transportation in OSCs. The surface morphology study and characterizations of PBDB-T-Ge:Y6 (1:1.2, w/w) blend film was performed using atomic force microscopy (AFM) and electron-scanning microscopy (SEM), as depicted in Figure 7. The biphasic domain of the PBDB-T-Ge:Y6 composite film can clearly be seen in the morphology study.

3.6. Structural Optimization and Density Functional Theory (DFT) Calculations

Density functional theory (DFT) calculations were conducted to provide further insights into the fundamental aspects of the molecular architecture of germanium-based comonomers. All calculations were conducted with the Gaussian16 software package. Optimized geometries of all compounds were obtained at the B3LYP level of theory with the 6-31G(d) basis set for all atoms. Previous work [39,40,41] showed that the geometries, and energetics of similar systems obtained using the B3LYP level of theory and 6-31G(d) basis set showed better agreement with experiments. Frequencies were calculated to ensure the absence of imaginary frequencies in the lowest-energy state. The HOMO and LUMO Hartree energies obtained from the calculation were converted to eV. The electrical transport properties of a molecule depend on the energy gap of HOMO and LUMO orbitals. The calculated frontier molecular orbital structures of the complexes are presented in Figure 8.
We observed that the electron density distributions at the HOMO are highly localized on the electron-rich moiety of BDTT-Ge (M1) in the ground state while, in the excited state, it is distributed at the LUMO and highly localized in the electron-deficiency moiety BDD (M2). The transition of the electron density distributions at the HOMO and LUMO were observed from all cases of -C, -Si and -Ge.
DFT calculations of Ge-based comonomer (triethyl-Ge) estimated the HOMO of −5.02 eV and LUMO of −2.36 eV with the band gap of 2.66 eV, while the band gaps for the counter-partner C-based comonomer (triethyl-C) and Si-based comonomer (triethyl-Si) were calculated to be 2.655 and 2.68 eV, respectively. The calculated electronic energy values of the HOMO, LUMO and band gap of the three co-monomers are summarized in Table 3. Clearly, the transition of charge density observed at LUMO and HOMO shows that the conjugated D-A comonomer can be a promising candidate for PSC, allowing for an efficient photo-induced intramolecular charge transfer (ICT) effect of the molecules.

4. Conclusions

A new 2D-conjugated donor–acceptor (D-A) copolymer (PBDB-T-Ge) was synthesized, where two reaction methods were compared to implement triethyl Ge: the conventional n-BuLi basis and Turbo–Grignard reactions. The Turbo–Grignard reaction was observed to be efficient in incorporating the group IV element to the side chain with a yield of 86%. The PBDB-T-Ge exhibited a LUMO of −3.64 eV, which was closer to that of Y6, an electron-acceptor. The gap between the LUMO values was calculated to be ca. 0.22 eV, which is sufficient to expect efficient photo-induced charge injection between the electron-donor and -acceptor in BHJ PSCs. PBDB-T-Ge and Y6 showed a complementary absorption, indicating a wide range of light-harvesting. DFT calculations on the comonomer verified that there were no significant changes in the band gap when the Ge sidegroup of the polymer replaces carbon- and silicon-base sidegroups.

Author Contributions

Conceptualization, W.H.A.; methodology, W.H.A. and D.T. and D.Y.; data curation, W.H.A. and S.M.I.; writing—original draft preparation, W.H.A.; writing—review and editing, Y.-G.K.; supervision, Y.-G.K.; funding acquisition, C.W. and Y.-G.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the DOD HBCU-MI Program (W911NF1810471), the DE NASA EPSCoR RID Seed Grant Program (80NSSC19M0087, 59602), and the NSF PREM program (2122158).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors would like to acknowledge Hai-Feng (Frank) Ji and his research group’s support for the NMR study. OSCAR (M. Moore) at DSU is acknowledged for supporting the morphology study. DFT calculations were supported by the COVID-19 High Performance Computing (HPC) Consortium project, and the authors used the Extreme Science and Engineering Discovery Environment (XSEDE), which was supported by NSF grant number ACI-1548562. Specifically, the authors acknowledge the Bridges-2 system, which was supported by NSF award number ACI-1928147 at the Pittsburgh Supercomputing Center (PSC) through the allocation of CHE210078 to S.M.I. This research was primarily/partially supported by NSF through the Delaware State University of Partnership in Research and Education in Materials, Division of Materials Research DMR. Additional support received from 2122158. The authors also acknowledge the use of facilities and instrumentation supported by NSF through the University of Delaware Materials Research Science and Engineering Center, DMR-2011824. The authors would like to thank T. Epps Group (Robert O’Dea) for the support of the GPC study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xu, Y.; Ji, Q.; Yin, L.; Zhang, N.; Liu, T.; Li, N.; He, X.; Wen, G.; Zhang, W.; Yu, L.; et al. Synergistic engineering of substituents and backbones on donor polymers: Toward terpolymer design of high-performance polymer solar cells. ACS Appl. Mater. Interfaces 2021, 13, 23993–24004. [Google Scholar] [CrossRef] [PubMed]
  2. Zhao, G.; He, Y.; Xu, Z.; Hou, J.; Zhang, M.; Min, J.; Chen, H.Y.; Ye, M.; Hong, Z.; Yang, Y.; et al. Effect of carbon chain length in the substituent of PCBM-like molecules on their photovoltaic properties. Adv. Funct. Mater. 2010, 20, 1480–1487. [Google Scholar] [CrossRef]
  3. Zhang, J.; Tan, H.S.; Guo, X.; Facchetti, A.; Yan, H. Material insights and challenges for non-fullerene organic solar cells based on small molecular acceptors. Nat. Energy 2018, 3, 720–731. [Google Scholar] [CrossRef]
  4. Wang, N.; Chen, Z.; Wei, W.; Jiang, Z. Fluorinated benzothiadiazole-based conjugated polymers for high-performance polymer solar cells without any processing additives or post-treatments. J. Am. Chem. Soc. 2013, 135, 17060–17068. [Google Scholar] [CrossRef]
  5. Zhou, R.; Jiang, Z.; Yang, C.; Yu, J.; Feng, J.; Adil, M.A.; Deng, D.; Zou, W.; Zhang, J.; Lu, K.; et al. All-small-molecule organic solar cells with over 14% efficiency by optimizing hierarchical morphologies. Nat. Commun. 2019, 10, 5393. [Google Scholar] [CrossRef]
  6. Lin, Y.; Firdaus, Y.; Nugraha, M.I.; Liu, F.; Karuthedath, S.; Emwas, A.H.; Zhang, W.; Seitkhan, A.; Neophytou, M.; Faber, H.; et al. 17.1% efficient single-junction organic solar cells enabled by n-type doping of the bulk-heterojunction. Adv. Sci. 2020, 7, 1903419. [Google Scholar] [CrossRef]
  7. Zhao, W.; Li, S.; Yao, H.; Zhang, S.; Zhang, Y.; Yang, B.; Hou, J. Molecular optimization enables over 13% efficiency in organic solar cells. J. Am. Chem. Soc. 2017, 139, 7148–7151. [Google Scholar] [CrossRef]
  8. Fan, B.; Zhong, W.; Ying, L.; Zhang, D.; Li, M.; Lin, Y.; Xia, R.; Liu, F.; Yip, H.L.; Li, N.; et al. Surpassing the 10% efficiency milestone for 1-cm2 all-polymer solar cells. Nat. Commun. 2019, 10, 4100. [Google Scholar] [CrossRef]
  9. Nagarjuna, P.; Bagui, A.; Hou, J.; Singh, S.P. New electron acceptor derived from fluorene: Synthesis and its photovoltaic properties. J. Phys. Chem. C 2016, 120, 13390–13397. [Google Scholar] [CrossRef]
  10. Singh, S.P.; Kumar, C.P.; Sharma, G.D.; Kurchania, R.; Roy, M.S. Synthesis of a modified PC70BM and its application as an electron acceptor with poly(3-hexylthiophene) as an electron donor for efficient bulk heterojunction solar cells. Adv. Funct. Mater. 2012, 22, 4087–4095. [Google Scholar] [CrossRef]
  11. Kuhlmann, J.C.; de Bruyn, P.; Bouwer, R.K.M.; Meetsma, A.; Blom, P.W.M.; Hummelen, J.C. Improving the compatibility of fullerene acceptors with fluorene-containing donor-polymers in organic photovoltaic devices. Chem. Commun. 2010, 46, 7232–7234. [Google Scholar] [CrossRef] [PubMed]
  12. Dai, S.; Zhao, F.; Zhang, Q.; Lau, T.K.; Li, T.; Liu, K.; Ling, Q.; Wang, C.; Lu, X.; You, W.; et al. Fused nonacyclic electron acceptors for efficient polymer solar cells. J. Am. Chem. Soc. 2017, 139, 1336–1343. [Google Scholar] [CrossRef] [PubMed]
  13. Yu, Y.; Jin, B.; Peng, R.; Fan, L.; Cai, L.; Fan, B.; Chu, S. Photovoltaic performance of PCBM analogs with different ester groups as acceptor in the polymer solar cells. Synth. Met. 2016, 212, 44–50. [Google Scholar] [CrossRef]
  14. Hwang, Y.J.; Courtright, B.A.E.; Ferreira, A.S.; Tolbert, S.H.; Jenekhe, S.A. 7.7% Efficient all-polymer solar cells. Adv. Mater. 2015, 27, 4578–4584. [Google Scholar] [CrossRef] [PubMed]
  15. Zhong, Y.; Trinh, M.T.; Chen, R.; Purdum, G.E.; Khlyabich, P.P.; Sezen, M.; Oh, S.; Zhu, H.; Fowler, B.; Zhang, B.; et al. Molecular helices as electron acceptors in high-performance bulk heterojunction solar cells. Nat. Commun. 2015, 6, 8242. [Google Scholar] [CrossRef]
  16. Kim, T.; Kim, J.H.; Kang, T.E.; Lee, C.; Kang, H.; Shin, M.; Wang, C.; Ma, B.; Jeong, U.; Kim, T.S.; et al. Flexible, highly efficient all-polymer solar cells. Nat. Commun. 2015, 6, 8547. [Google Scholar] [CrossRef]
  17. Zhang, Y.; Cai, G.; Li, Y.; Zhang, Z.; Li, T.; Zuo, X.; Lu, X.; Lin, Y. An electron acceptor analogue for lowering trap density in organic solar cells. Adv. Mater. 2021, 33, 2008134. [Google Scholar] [CrossRef]
  18. Li, Y.; Liu, X.; Wu, F.P.; Zhou, Y.; Jiang, Z.Q.; Song, B.; Xia, Y.; Zhang, Z.G.; Gao, F.; Inganäs, O.; et al. Non-fullerene acceptor with low energy loss and high external quantum efficiency: Towards high performance polymer solar cells. J. Mater. Chem. A 2016, 4, 5890–5897. [Google Scholar] [CrossRef]
  19. Lin, Y.; Zhao, F.; He, Q.; Huo, L.; Wu, Y.; Parker, T.C.; Ma, W.; Sun, Y.; Wang, C.; Zhu, D.; et al. High-performance electron acceptor with thienyl side chains for organic photovoltaics. J. Am. Chem. Soc. 2016, 138, 4955–4961. [Google Scholar] [CrossRef]
  20. Duan, Y.; Xu, X.; Li, Y.; Li, Z.; Peng, Q. Chalcogen-atom-annulated perylene diimide trimers for highly efficient nonfullerene polymer solar cells. Macromol. Rapid Commun. 2017, 38, 1700405. [Google Scholar] [CrossRef]
  21. Guo, Y.; Li, Y.; Awartani, O.; Zhao, J.; Han, H.; Ade, H.; Zhao, D.; Yan, H. A Vinylene-bridged perylenediimide-based polymeric acceptor enabling efficient all-polymer solar cells processed under ambient conditions. Adv. Mater. 2016, 28, 8483–8489. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, Z.; Feng, L.; Xu, S.; Liu, Y.; Peng, H.; Zhang, Z.G.; Li, Y.; Zou, Y. A new electron acceptor with meta-alkoxyphenyl side chain for fullerene-free polymer solar cells with 9.3% efficiency. Adv. Sci. 2017, 4, 1700152. [Google Scholar] [CrossRef] [PubMed]
  23. Qiu, N.; Zhang, H.; Wan, X.; Li, C.; Ke, X.; Feng, H.; Kan, B.; Zhang, H.; Zhang, Q.; Lu, Y.; et al. A new nonfullerene electron acceptor with a ladder type backbone for high-performance organic solar cells. Adv. Mater. 2017, 29, 1604964. [Google Scholar] [CrossRef] [PubMed]
  24. Zhao, F.; Dai, S.; Wu, Y.; Zhang, Q.; Wang, J.; Jiang, L.; Ling, Q.; Wei, Z.; Ma, W.; You, W.; et al. Single-junction binary-blend nonfullerene polymer solar cells with 12.1% efficiency. Adv. Mater. 2017, 29, 1700144. [Google Scholar] [CrossRef]
  25. Bin, H.; Zhang, Z.G.; Gao, L.; Chen, S.; Zhong, L.; Xue, L.; Yang, C.; Li, Y. Non-fullerene polymer solar cells based on alkylthio and fluorine substituted 2D-conjugated polymers reach 9.5% efficiency. J. Am. Chem. Soc. 2016, 138, 4657–4664. [Google Scholar] [CrossRef]
  26. Guo, B.; Li, W.; Guo, X.; Meng, X.; Ma, W.; Zhang, M.; Li, Y. High efficiency nonfullerene polymer solar cells with thick active layer and large area. Adv. Mater. 2017, 29, 1702291. [Google Scholar] [CrossRef]
  27. Chen, S.; Liu, Y.; Zhang, L.; Chow, P.C.Y.; Wang, Z.; Zhang, G.; Ma, W.; Yan, H. A wide-bandgap donor polymer for highly efficient non-fullerene organic solar cells with a small voltage loss. J. Am. Chem. Soc. 2017, 139, 6298–6301. [Google Scholar] [CrossRef]
  28. Wang, J.; Wang, W.; Wang, X.; Wu, Y.; Zhang, Q.; Yan, C.; Ma, W.; You, W.; Zhan, X. Enhancing performance of nonfullerene acceptors via side-chain conjugation strategy. Adv. Mater. 2017, 29, 1702125. [Google Scholar] [CrossRef]
  29. Liu, J.; Chen, S.; Qian, D.; Gautam, B.; Yang, G.; Zhao, J.; Bergqvist, J.; Zhang, F.; Ma, W.; Ade, H.; et al. Fast charge separation in a non-fullerene organic solar cell with a small driving force. Nat. Energy 2016, 1, 16089. [Google Scholar] [CrossRef]
  30. Ma, R.; Liu, T.; Luo, Z.; Guo, Q.; Xiao, Y.; Chen, Y.; Li, X.; Luo, S.; Lu, X.; Zhang, M.; et al. Improving open-circuit voltage by a chlorinated polymer donor endows binary organic solar cells efficiencies over 17%. Sci. China Chem. 2020, 63, 325–330. [Google Scholar] [CrossRef]
  31. Weng, K.; Ye, L.; Zhu, L.; Xu, J.; Zhou, J.; Feng, X.; Lu, G.; Tan, S.; Liu, F.; Sun, Y. Optimized active layer morphology toward efficient and polymer batch insensitive organic solar cells. Nat. Commun. 2020, 11, 2855–2863. [Google Scholar] [CrossRef] [PubMed]
  32. Zhu, L.; Zhang, M.; Zhou, G.; Hao, T.; Xu, J.; Wang, J.; Qiu, C.; Prine, N.; Ali, J.; Feng, W.; et al. Efficient organic solar cell with 16.88% efficiency enabled by refined acceptor crystallization and morphology with improved charge transfer and transport properties. Adv. Energy Mater. 2020, 10, 1904234. [Google Scholar] [CrossRef]
  33. Chen, S.; Lee, S.M.; Xu, J.; Lee, J.; Lee, K.C.; Hou, T.; Yang, Y.; Jeong, M.; Lee, B.; Cho, Y.; et al. Ultrafast channel II process induced by a 3-D texture with enhanced acceptor order ranges for high-performance non-fullerene polymer solar cells. Energy Environ. Sci. 2018, 11, 2569–2580. [Google Scholar] [CrossRef]
  34. Bin, H.; Gao, L.; Zhang, Z.G.; Yang, Y.; Zhang, Y.; Zhang, C.; Chen, S.; Xue, L.; Yang, C.; Xiao, M.; et al. 11.4% Efficiency non-fullerene polymer solar cells with trialkylsilyl substituted 2D-conjugated polymer as donor. Nat. Commun. 2016, 7, 13651. [Google Scholar] [CrossRef]
  35. Sherborne, G.J.; Gevondian, A.G.; Funes-Ardoiz, I.; Dahiya, A.; Fricke, C.; Schoenebeck, F. Modular and selective arylation of aryl germanes (C−GeEt3) over C−Bpin, C−SiR3 and halogens enabled by light-activated gold catalysis. Angew. Chem. Int. Ed. 2020, 59, 15543–15548. [Google Scholar] [CrossRef]
  36. Sharma, S.; Soni, R.; Kurungot, S.; Asha, S.K. Rylene diimide-based alternate and random copolymers for flexible supercapacitor electrode materials with exceptional stability and highpower density. J. Phys. Chem. C 2019, 123, 2084–2093. [Google Scholar] [CrossRef]
  37. Carsten, B.; He, F.; Son, H.J.; Xu, T.; Yu, L. Stille polycondensation for synthesis of functional materials. Chem. Rev. 2011, 111, 1493–1528. [Google Scholar] [CrossRef]
  38. Qian, D.; Ye, L.; Zhang, M.; Liang, Y.; Li, L.; Huang, Y.; Guo, X.; Zhang, S.; Tan, Z.; Hou, J. Design, application, and morphology study of a new photovoltaic polymer with strong aggregation in solution state. Macromolecules 2012, 45, 9611–9617. [Google Scholar] [CrossRef]
  39. Al-Busaidi, I.J.; Haque, A.; Al-Balushi, R.A.; Rather, J.A.; Munam, A.; Ilmi, R.; Raithby, P.R.; Zhang, Y.; Fu, Y.; Xie, Z.; et al. Synthesis, characterization, and optoelectronic properties of phenothiazine-based organic co-polyynes. New J. Chem. 2021, 45, 15082–15095. [Google Scholar] [CrossRef]
  40. Jayapal, M.; Haque, A.; Al-Busaidi, I.J.; Al-Rasbi, N.; Al-Suti, M.K.; Khan, M.S.; Al-Balushi, R.; Islam, S.M.; Xin, C.; Wu, W.; et al. Dicopper(I) Complexes Incorporating Acetylide-Functionalized Pyridinyl-Based Ligands: Synthesis, Structural, and Photovoltaic Studies. Inorg. Chem. 2018, 57, 12113–12124. [Google Scholar] [CrossRef]
  41. Chowdhury, F.I.; Islam, J.; Arof, A.K.; Khandaker, M.U.; Zabed, H.M.; Khalil, I.; Rahman, M.R.; Islam, S.M.; Karim, M.R.; Uddin, J. Electrocatalytic and structural properties and computational calculation of PAN-EC-PC-TPAI-I 2 gel polymer electrolytes for dye sensitized solar cell application. RSC Adv. 2021, 11, 22937–22950. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Molecular structures of PBDB-T-C (left), PBDB-T-Si (center), and PBDB-T-Ge (right).
Figure 1. Molecular structures of PBDB-T-C (left), PBDB-T-Si (center), and PBDB-T-Ge (right).
Polymers 15 02429 g001
Scheme 1. Schematic illustration of synthetic route of PBDB-T-Ge.
Scheme 1. Schematic illustration of synthetic route of PBDB-T-Ge.
Polymers 15 02429 sch001
Figure 2. 1H NMR for triethyl(thiophen-2-yl) germane (2).
Figure 2. 1H NMR for triethyl(thiophen-2-yl) germane (2).
Polymers 15 02429 g002
Figure 3. UV-Vis absorption of BDTT-Ge (M1), BDD (M2) and PBDB-T-Ge.
Figure 3. UV-Vis absorption of BDTT-Ge (M1), BDD (M2) and PBDB-T-Ge.
Polymers 15 02429 g003
Figure 4. UV-Vis absorption of PBDB-T-Ge and Y6.
Figure 4. UV-Vis absorption of PBDB-T-Ge and Y6.
Polymers 15 02429 g004
Figure 5. Photoluminescence (PL)-quenching profile of PBDB-T-Ge with PC60BM in chloroform solution. The inset represents the fluorescence light emission of PBDB-T-Ge in chloroform under UV light. The emission was observed to decrease as increasing PC60BM in the polymer solution.
Figure 5. Photoluminescence (PL)-quenching profile of PBDB-T-Ge with PC60BM in chloroform solution. The inset represents the fluorescence light emission of PBDB-T-Ge in chloroform under UV light. The emission was observed to decrease as increasing PC60BM in the polymer solution.
Polymers 15 02429 g005
Figure 6. (a) Cyclic voltammogram of PBDB-T-Ge polymer film on a glassy carbon electrode was measured in 0.1 M Et4NBF4 acetonitrile solution at a scan rate of 10 mV/s, with the cyclic voltammogram of ferrocene/ferrocenium (Fc/Fc+) pair used as an internal reference. (b) A diagram indicating the energy levels of polymers PBDB-T-C, PBDB-T-Si and PBDB-T-Ge. The energy level of PBDB-T-Ge is compared with that of Y6. The energy levels of PBDB-T-C and PBDB-T-Si were obtained from references [33,38].
Figure 6. (a) Cyclic voltammogram of PBDB-T-Ge polymer film on a glassy carbon electrode was measured in 0.1 M Et4NBF4 acetonitrile solution at a scan rate of 10 mV/s, with the cyclic voltammogram of ferrocene/ferrocenium (Fc/Fc+) pair used as an internal reference. (b) A diagram indicating the energy levels of polymers PBDB-T-C, PBDB-T-Si and PBDB-T-Ge. The energy level of PBDB-T-Ge is compared with that of Y6. The energy levels of PBDB-T-C and PBDB-T-Si were obtained from references [33,38].
Polymers 15 02429 g006
Figure 7. Surface morphology of (a) AFM 3D topography images (um in x-y unit) and (b) SEM image of the PBDB-T-Ge:Y6 composite film.
Figure 7. Surface morphology of (a) AFM 3D topography images (um in x-y unit) and (b) SEM image of the PBDB-T-Ge:Y6 composite film.
Polymers 15 02429 g007
Figure 8. (a) Optimized geometry (left) and HOMO and LUMO electron-density plots (right) of triethyl-Ge via DFT calculations of triethyl-Ge; (b) HOMO and LUMO electron-density plots of triethyl-C and triethyl-Si via DFT calculations. Note that the sidegroup effect of triethyl-Ge (3rd) of triethyl-Ge is compared with triethyl-C (left in (b)) and -Si (right in (b)).
Figure 8. (a) Optimized geometry (left) and HOMO and LUMO electron-density plots (right) of triethyl-Ge via DFT calculations of triethyl-Ge; (b) HOMO and LUMO electron-density plots of triethyl-C and triethyl-Si via DFT calculations. Note that the sidegroup effect of triethyl-Ge (3rd) of triethyl-Ge is compared with triethyl-C (left in (b)) and -Si (right in (b)).
Polymers 15 02429 g008
Table 1. UV-Vis and emission characteristics of PBDB-T-Ge.
Table 1. UV-Vis and emission characteristics of PBDB-T-Ge.
MaterialsλmaxEx (nm)λonset (nm)λmaxEm
(nm)
SolutionFilmSolutionFilm
PBDB-T-Ge477507601666615
Table 2. Electronic Energy Properties of PBDB-T-C, PBDB-T-Si, and PBDB-T-Ge.
Table 2. Electronic Energy Properties of PBDB-T-C, PBDB-T-Si, and PBDB-T-Ge.
MaterialsHOMO
(eV)
LUMO
(eV)
Egopt
(eV) a
Egcv
(eV) b
PBDB-T-C c−5.23 −3.18-2.05
PBDB-T-Si c−5.45 −3.581.81-
PBDB-T-Ge−5.45 −3.641.861.81
Y6−5.65 −3.861.341.79
a Band gaps were calculated from the onset of the film absorption. b Measured by cyclic voltammetry. c Data obtained from references [33,38].
Table 3. DFT Calculation Results of Comonomers of PBDB-T.
Table 3. DFT Calculation Results of Comonomers of PBDB-T.
Side Groups aHOMO
(eV)
LUMO
(eV)
EgDFT
(eV) a
Triethyl-C−5.01−2.352.65
Triethyl-Si−5.06−2.382.68
Triethyl-Ge−5.02−2.362.66
a The side group, triethyl-Ge of PBDB-T-Ge was replaced with triethyl-C and triethyl-Si for the DFT calculations.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Abousamra, W.H.; Thomas, D.; Yang, D.; Islam, S.M.; Winstead, C.; Kim, Y.-G. Synthesis and Characterization of the Donor-Acceptor Conjugated Polymer PBDB-T Implementing Group IV Element Germanium. Polymers 2023, 15, 2429. https://doi.org/10.3390/polym15112429

AMA Style

Abousamra WH, Thomas D, Yang D, Islam SM, Winstead C, Kim Y-G. Synthesis and Characterization of the Donor-Acceptor Conjugated Polymer PBDB-T Implementing Group IV Element Germanium. Polymers. 2023; 15(11):2429. https://doi.org/10.3390/polym15112429

Chicago/Turabian Style

Abousamra, Wafaa H., Destinee Thomas, Dan Yang, Shahidul M. Islam, Cherese Winstead, and Young-Gi Kim. 2023. "Synthesis and Characterization of the Donor-Acceptor Conjugated Polymer PBDB-T Implementing Group IV Element Germanium" Polymers 15, no. 11: 2429. https://doi.org/10.3390/polym15112429

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop