Axially phenoxylated aluminum phthalocyanines and their application in organic photovoltaic cells

Hasan Rabouia, Mohammad AL-Amara, Ahmed I. Abdelrahmana and Timothy P. Bender*abc
aUniversity of Toronto, Dept. of Chemical Engineering & Applied Chemistry, 200 College Street, Toronto, Ontario M5S 3E5, Canada. E-mail: tim.bender@utoronto.ca
bUniversity of Toronto, Dept. of Materials Science and Engineering, 200 College Street, Toronto, Ontario M5S 3E5, Canada
cUniversity of Toronto, Dept. of Chemistry, 80 St. George Street, Toronto, Ontario M5S 3H6, Canada

Received 19th March 2015 , Accepted 7th May 2015

First published on 7th May 2015


Abstract

Novel axially phenoxylated aluminium phthalocyanines, pentafluorophenoxy aluminium phthalocyanine (F5PhO-AlPc) and p-nitrophenoxy aluminium phthalocyanine (NO2PhO-AlPc), were synthesized through a one-step reaction starting from the commonly used photoactive material, chloro aluminium phthalocyanine (Cl-AlPc), and the respective phenols. Reactions with other phenols did not yield corresponding AlPc derivatives. Optical, electrical, and thermal analyses were carried out on F5PhO-AlPc and NO2PhO-AlPc using UV-Vis measurements (solution and thin-film), cyclic voltammetry (CV), differential-pulse voltammetry (DPV), and thermogravimetric analysis (TGA). A simple thermodynamic model was used to explain the lack of reaction when Cl-AlPc was treated a variety of alkylated phenols. We noted a side reaction producing fluoro aluminium phthalocyanine (F-AlPc) when the synthesis of F5PhO-AlPc was attempted in DMF. The model also explains this observation. F5PhO-AlPc and NO2PhO-AlPc were integrated into organic photovoltaic devices (OPVs) both as electron-donating and as electron-accepting materials. The phenoxy-AlPcs enable an enhancement of the open-circuit voltage (VOC) of the OPVs when applied as either an electron donor or as an electron acceptor compared to Cl-AlPc. The results within the OPV, specifically the increased VOC, are consistent with the steric shielding effect seen in other OPVs.


Introduction

Phthalocyanines (Pcs) are a family of molecules initially discovered in the early 20th century and have since been industrially produced and used extensively in the pigment industry.1,2 Pcs possess certain characteristics which enable their use within the pigment industry including high thermal- and photo-stability, ease of synthesis through a one-step reaction (generally) and high molar extinction coefficients. More recently Pcs have demonstrated the ability to move electrical charge and be applied as organic semiconductors in a variety of organic electronic applications including organic photoreceptors,3–5 organic photovoltaics (OPV)6–11 and organic light emitting diodes (OLED).12,13 Pcs have a tight connection to OPVs as one of the first functioning OPVs reported in the literature used copper phthalocyanine (CuPc) as an electrondonor material with a power conversion efficiency just under 1%.14 Since then, many OPV devices have been fabricated using different Pc derivatives whereby Pcs have most commonly been applied as electron donating mateirals6,15,16 or as light sensitizers.17–20 Xue et al. achieved efficiencies up to 5% using Pcs in the light-absorbing layer.21 There is, however, limited research on using Pcs as electron acceptor materials in OPVs. Very recently, the group of Jones has demonstrated that Cl-AlPc can function as an electron acceptor material when paired with pentacene in an OPV.22

Research into Pcs whereby the intended application is within an OPV has divided into two discreet areas: developing new Pcs through chemical modification9,23–25 or altering the device fabrication process and/or configuration. Each avenue is generally targeting OPVs with higher photo conversion efficiencies.10,11,26–28 However, achieving better basic characteristics such as enhanced OPV open circuit voltage (VOC) through chemical modification is also highly desirable.

Regarding chemical modification in general, there has been extensive research over past decades looking at the chemistry of Pcs through changing of the central atom,1 peripheral substitution29–33 or through axial substitution.34–38 Nearly all elements in the periodic table (95 elements) have been successfully incorporated into Pcs.1 Some of the most common central atoms are divalent such as Cu, Zn, and Mg. In these divalent cases, tuning the molecular properties is only possible through peripheral substitution and has been shown to affect the solubility,29,30 stability,29 optical properties31 and electrical properties of the resulting Pcs.32,33

Trivalent (e.g. Al) and tetravalent (e.g. Si) metal phthalocyanines (AlPc and SiPc respectively) have the unique ability to form axial bonds with different chemical functionalities that can be potentially used to influence the properties of the resulting Pc. However, limited reports are available detailing the chemistry of the axial position of trivalent and tetravalent Pcs and the impact that axial modifications have on the properties of the Pcs. The majority of the reports that are available detail the chemistry of axial substitution of silicon phthalocyanine (SiPc)34 and boron subphthalocyanine (BsubPc)35–40 and include μ-oxo41–44 and polymeric Pc derivatives.45–47

Our group has shown the axial substitution of BsubPcs with a pentafluoro phenoxy fragment can alter the frontier orbital energies and change the OPV performance characteristics.48 We have also recently shown that axial phenoxylation can change SiPc from being largely electronically inactive to something that has potential as an electron acceptor in a fullerene free OPV.49 Finally, we have demonstrated that changing the axial chloride of Cl-AlPc to a fluoride (F-AlPc) results in major changes in the fundamental properties of the AlPc including its solid state arrangement.50

Beside our recently study of F-AlPc, the axial chemistry of AlPc has been relatively understudied. Kraus and Louw,51 Brasseur et al.,52 and Sakai et al.53 have shown that silver based reagents can be used to axially phenoxylate AlPcs although in only one case was the yield reported and in no cases was the conversion reported.

In this paper, we outline our independent investigation into the axial phenoxylation of chloro aluminum phthalocyanine (Cl-AlPc) with a specific focus on finding a general, easy, inexpensive and high yielding method for phenoxylation whereby high conversion is achieved. High conversion is desirable to simplify the purification process of the material and to minimize the potentiality of impurities being present within the final materials as they are incorporated into organic electronic devices. The investigation of the resulting derivatives in planar heterojunction (PHJ) organic photovoltaic cells (OPVs) was also undertaken. Despite multiple attempts with variety of phenols, only two derivatives, pentafluorophenoxy aluminum phthalocyanine (F5PhO-AlPc) and p-nitrophenoxy aluminum phthalocyanine (NO2PhO-AlPc) were successfully synthesized and characterized. Thermodynamic limitations on axial phenoxylation of a range of metal Pcs were calculated. Effects on the optical properties, redox potentials, solid-state packing, and stability were observed. OPVs were fabricated employing F5PhO-AlPc and NO2PhO-AlPc both as an electron donor and as an electron acceptor. OPV characteristics were compared against Cl-AlPc baseline devices.

Experimental section

Materials

All ACS grade solvents (except nitrobenzene and quinoline) were purchased from Caledon Labs (Caledon, Ontario, Canada) and used without further purification unless otherwise stated. Phthalonitrile was purchased from TCI America (Portland, Oregon, USA) and used as received. Aluminum chloride, p-nitrophenol, m-cresol, 3-pentadecylphenol, 2-sec-butylphenol, 4-tert-octylphenol, nitrobenzene, quinoline, decamethyl ferrocene and tetrabutylammonium perchlorate were purchased from Sigma-Aldrich Chemical Company (Mississauga, Ontario, Canada) and used as received. Pentafluoro phenol was purchased from Oakwood Products, Inc. (Columbia, South Carolina, USA) and used as received. Bathocuproine (BCP 99.99%) from Sigma-Aldrich, Ag (99.99%) from R. D. Mathis Company, Fullerene (C60) from SES Research and α-sexithiophene (α-6T) from Sigma-Aldrich were all used as received.

General methods

All reactions were performed under an atmosphere of nitrogen gas using oven-dried glassware. Low-resolution mass spectra (LRMS) and high-resolution mass spectra (HRMS) were determined using Waters GC Premier using time-of-flight spectrometer with electron ionization source (TOF-EI MS). X-ray diffraction data were collected on a Bruker Kappa APEX-DUO diffractometer using a Copper ImuS (microsource) tube with multi-layer optics (F5PhO-AlPc) or a molybdenum sealed tube with Triumph momochromator (NO2PhO-AlPc) and were measured using a combination of f scans and w scans. The data were processed using APEX2 and SAINT.54 Absorption corrections were carried out using SADABS.54 The structure was solved and refined using SHELXT & SHELXL-2103 (ref. 55) for full-matrix least-squares refinement that was based on F2. Ultraviolet-visible (UV-Vis) absorption spectra were acquired on a PerkinElmer Lambda 1050 UV/VIS/NIR spectrometer using a PerkinElmer quartz cuvette with a 10 mm path length. Cyclic voltammetry and differential-pulse voltammetry were carried out using a Bioanalytical Systems C3 electrochemical workstation. The working electrode was a 1 mm glassy carbon disk, the counter electrode was a platinum wire, and the reference electrode was Ag/AgCl saturated salt solution. Spec-grade solvents were purged with nitrogen gas at room temperature prior to their use. The scanning rate applied throughout both experiments is 100 mV s−1. Tetrabutyl ammonium perchlorate (0.1 M) was used as the supporting electrolyte and decamethyl ferrocene as an internal reference. All half-wave potentials were corrected to the half-wave reduction potential (E1/2,red) of decamethyl ferrocene. Thermogravimetric analysis was run using TA Instruments Q50 under an atmosphere of nitrogen and a heating rate of 5 °C min−1. Soxhlet extractions were performed using Whatman® single thickness cellulose extraction thimbles.

Synthetic methods

Chloro aluminum phthalocyanine (Cl-AlPc). Chloro aluminum phthalocyanine (Cl-AlPc) was synthesized as described by Santerre et al.56 20 g (0.16 mol) of phthalonitrile was added to a solvent-mixture of quinolone (20 mL) and nitrobenzene (32 mL) in a 250 mL round-bottom flask. The solution was cooled to 0 °C in an ice-bath and 5.72 g (0.0439 mol) of AlCl3 was added to the mixture. The reaction was brought up to reflux at 210 °C and left stirring for 4 hours. The resulting blue-greenish solution was cooled to room temperature. The crude product was filtered using a glass-frit filter and washed thoroughly with acetone and acetonitrile. A Soxhlet extraction with acetone was performed overnight on the resulting cake. After the product was dried overnight in the vacuum-oven, 11.2 g (49.9% yield) of blue powder was collected. TOF-EI MS (Fig. S1) m/z (%): 539.1 [MPc+] (30), 540.1[MPc+] (10), 574.1 (100), 575.1 (20). The product was then purified through conventional temperature-gradient train sublimation apparatus (as describe elsewhere57). Anal. Calcd for C32H16N8AlCl: C, 66.85; H, 2.81; N, 19.49. Found: C, 66.57; H, 3.38; N, 19.37%.
Pentafluorophenoxy aluminum phthalocyanine (F5PhO-AlPc). Pentafluorophenoxy aluminum phthalocyanine (F5PhO-AlPc) was synthesized by an adapted method from the synthesis of F5BsubPc.58 1.0 g (1.7 mmol) of Cl-AlPc was mixed with 5 molar excess of pentafluoro phenol (1.7 g) in 20 mL of toluene. The reaction mixture was heated and kept under reflux for 24 hours. After the mixture was allowed to cool down, toluene was removed using rotary evaporation. The powder was then washed successively with toluene. The resulting blue powder was dried in a vacuum oven overnight. 1.2 g (∼100% yield) of the product was collected. TOF-EI (Fig. S2) m/z (%): 539.1 [MPc+] (100), 540.1[MPc+] (20), 722.1 (100), 723.1 (20). Further purification of the product (0.56 g) took place using the temperature gradient train sublimation apparatus (as describe elsewhere57) yielding 49.3% of pure product. TOF-EI MS m/z (%): 539.1 [MPc+] (100), 540.1[MPc+] (20), 722.1 (100), 723.1 (20). Anal. calcd for C38H16N8AlOF5: C, 63.17; H, 2.23; N, 15.51. Found: C, 63.26; H, 2.76; N, 15.59%.
p-Nitrophenoxy aluminum phthalocyanine (NO2PhO-AlPc). p-Nitrophenoxy aluminum phthalocyanine (NO2PhO-AlPc) was also, synthesized by an adapted method from the synthesis of F5BsubPc.58 0.5 g (1.35 mmol) of Cl-AlPc was mixed with 5 molar excess of p-nitrophenol (0.6 g) in 10 mL of chlorobenzene. The reaction mixture was heated and kept under reflux for 24 hours. After the mixture was allowed to cool down, chlorobenzene was removed using rotary evaporation. The powder was then washed with ethyl ether. The resulting blue powder was dried in a vacuum oven overnight. 0.5 g (84.9% yield) of the product was collected. TOF-EI MS (Fig. S4) m/z (%): 539.1 [MPc+] (100), 540.1[MPc+] (20), 677.2 (50), 678.2 (10). The reaction was also run in DMF producing similar results. The product was then taken into the train sublimation apparatus at 450 °C; the yield of the sublimed product, however, was low (∼10%). Anal. calcd for C38H16N9AlO3: C, 67.36; H, 2.98; N, 18.60. Found: C, 66.78; H, 2.29; N, 18.57%.

Device fabrication

All OPVs device were fabricated on commercially available patterned ITO substrates (Thin Film Device Inc. (TFD)) with 145 nm ITO thickness and a sheet resistance of 15 Ω □−1. The ITO surfaces were ultrasonically cleaned in detergent, distilled water, acetone, methanol, and finally dried in flowing nitrogen. Subsequently, ITO was treated by plasma for 5 min to remove carbon residues. After plasma treatment, a CHEMAT Technology KW-A4 spin coater was used to spin a layer of PEDOT:PSS (Clevios™ P VP AI 4083) at 500 rpm for 10 s and 4000 rpm for 30 s on the ITO electrode anode. The ITO with the PEDOT:PSS coating was annealed in a at 110 °C for 15 min. The device structure grown by thermal evaporation with a base pressures of 10−8 torr consists ITO/PEDOT:PSS/Active layer/BCP(8 nm)/Ag(80 nm), where active layer and their thickness can be found in Table 2. Between the deposition of the BCP layer and the Ag cathode, devices were transferred from the vacuum chamber to nitrogen atmosphere glove box through a gate, without a vacuum break, to attach a shadow mask consisting of 5 active pixel area of 0.2 cm2 openings then transferred back to the vacuum chamber. Layers thicknesses and deposition rates were monitored using a quartz crystal microbalance (QCM). After the Ag layers were deposited, the OPV devices were transferred directly from the vacuum chamber back to the nitrogen atmosphere glove box. Silver paste (PELCO® Conductive Silver 187) was applied to the Ag and to the ITO electrodes to enhance electrical contact and left to cure for 30 minutes before testing. A 300 W ozone free xenon arc lamp with an air Mass 1.5 Global filter, fed through a Cornerstone™ 260 1/4 m monochromator and then into the glove box by way of a single branch liquid light guide was utilized as a solar simulator. The illumination levels for the test cell were kept at 100 mW cm−2. A UV-silicon photo-detector was used to calibrate the measurements. The current densities versus voltage (JV) characteristics and external quantum efficiency were recorded using a Keithley 2401 Low Voltage sourcemeter controlled by a custom LabView program and a Newport Optical Power Meter 2936-R controlled by TracQ Basic software, respectively, in the nitrogen atmosphere glove box.
Table 1 Oxidation and reduction potentials using CV and DPV (in dichloromethane), electrical band gap (EEG), optical band gap (EOG), and the estimated frontier energy levels of AlPc derivatives
  Eox (V) Ered (V) EEG (eV) EOG (eV) HOMO (eV) LUMO (eV)
Cl-AlPc   1.40 5.3 (ref. 59) 4.0 (ref. 59)
F5PhO-AlPc 0.70 −0.71 1.70 1.51 5.52 3.83
NO2PhO-AlPc 0.60 −0.72 1.58 1.55 5.40 3.82


Results and discussion

We began this study by surveying the reaction of chloro aluminum phthalocyanine (Cl-AlPc) with a variety of phenols by adopting the simple method our laboratory has used for phenoxylation of chloro boron subphthalocyanine (Cl-BsubPc).58 This approach is high yielding for BsubPcs and involves nothing more than simple heating in solution with a molar excess of the appropriate phenol. Phenols that we tried included several alkylated phenols (m-cresol, 2-sec-butylphenol, 4-tert-octylphenol, and 3-pentadecylphenol) and two more ‘acidic’ phenols (pentafluoro phenol and 4-nitrophenol, Scheme 1). Compared against the reaction of Cl-BsubPc, conversion of Cl-AlPc to the corresponding phenoxylated derivatives was low and impractical when alkylated phenols were used. At low molar excesses of alkylated phenol (5 equiv) little to no conversion of Cl-AlPc was observed. Using large molar excesses (∼100–150 equiv) low conversion was at times observed (50% maximum) but was not reproducible. These observations are in line with that of Owen and Kenney.59 We did not consider the use of the methods described by Kraus and Louw,51 Brasseur et al.52 or Sakai et al.53 as all use expensive silver salts to activate the Al–Cl bond to reaction.
image file: c5ra04919a-s1.tif
Scheme 1 Phenoxylation reaction of Cl-AlPc. Where conversion to the phenoxylated AlPc was limited, a N/a is placed in the designation cell. Changing bonds are highlighted in red (breaking) and green (forming).

The only reproducible syntheses were found to be where the ‘acidic’ phenols pentafluoro phenol (F5PhOH) and 4-nitrophenol (NO2PhOH) were reacted with Cl-AlPc. F5PhO-AlPc was successfully synthesized using toluene as a solvent and NO2PhO-AlPc with chlorobenzene. The general workup of the reaction involved rotary evaporating the crude solution to dryness. The powder was then washed with an appropriate organic solvent and dried in a vacuum-oven overnight. Both F5PhOH and NO2PhOH produced high yields (>80%) of the respective AlPc derivatives. Each product was then purified by train sublimation.57

The identity of each phenoxylated AlPc was first confirmed using LR-MS and HR-MS (Fig. S2 and S3). Additionally, the identity of F5PhO-AlPc was confirmed by X-ray diffraction of solvated single crystals grown by solvent diffusion (heptane into a dilute solution of toluene; Fig. S4; Tables S1–S6). Given the crystals were solvated a discussion of the solid-state arrangement, relevant to those interested in solid-state applications such as OPV is not appropriate. Conversely, we were able to obtain single crystals of NO2PhO-AlPc grown by sublimation an analogous process to that which is used for fabrication of OPV devices (Fig. S5–S6, Tables S7–S12,). It was found that the solid-state packing of NO2PhO-AlPc is extremely dense (Fig. S6) with π–π interactions between the AlPc chromaphores at distances less than 3.8 Å in all three crystallographic axis; a desirable characteristic for an organic electronic material. The obvious point of comparison would be the solid-state arrangement of Cl-AlPc. While there is a crystal structure of Cl-AlPc report by Wynne,42 as noted there is significant disorder within the unit cell of the single crystal. The absolute positions of two Cl-AlPc molecules were determined but the position(s) of the remaining one or two molecules could not be determined.42 Therefore without a point of comparison, a more in depth discussion of the solid-state arrangement of NO2PhO-AlPc is irrelevant. We have made several attempts in our laboratory to obtain single crystals of Cl-AlPc without success.

Determination of the NMR spectra for both F5PhO-AlPc and NO2PhO-AlPc was not possible due to their limited solubility (common for Pcs). While each did have limited solubility in DMSO, DCM and toluene each had only enough solubility in DMSO to yield an NMR spectrum. NMR analysis in DSMO-d6 showed that each reacted with residual water to produce the same hydrolyzed product (presumably HO-AlPc). Final purity of each material was therefore confirmed using elemental analysis (see Experimental) prior to OPV device integration.

Given the polarity of the acidic phenols, we also tried the corresponding reactions in DMF. However, in DMF the reaction of pentafluoro phenol and Cl-AlPc yielded exclusively F-AlPc as confirmed by MS (Fig. S7). Shoute et al. has studied the defluoration of pentafluoro phenol and formation of fluoride ions in aqueous solutions.60 We hypothesize that the residual water present in reagent grade DMF (and not in reagent grade toluene) allowed for the generation of fluoride ions and their subsequent reaction with Cl-AlPc to produce F-AlPc. NO2PhO-AlPc could also be synthesized in toluene or DMF as well as chlorobenzene as the solvent without issues.

Thermodynamic limitations

As outlined above, only two out of the six phenols successfully formed phenoxylated AlPcs whereas multiple examples of phenoxylation of SiPcs34,49 and BsubPcs35–40 using similar processes are known. In an effort to explain these observations, we turned to an estimation of the enthalpy of the phenoxylation reaction (ΔHR) for different MPcs. The phenoxylation reaction was assumed to follow Scheme 1 and we have made no assumptions regarding the mechanism by which the reaction occurs. ΔHR was simply estimated using bond dissociation enthalpies of the bonds involved in the phenoxylation reaction for several MPcs (eqn (1)). To a first approximation, ΔHR is assumed to depend only on the bonds being broken and formed during the reaction and not being dependent on the molecule as a whole.61
 
ΔHR = (ΔHH–X + ΔHO–M) − (ΔHX–M + ΔHO–H) (1)
ΔHA–B is the enthalpy of formation of the bond between A and B. If we assume that the entropy of the products and the reactants are similar and that the changes in entropy can be neglected (since the number of moles throughout the reaction is constant and since the reaction is run in solution without a phase change) then the sign and magnitude of ΔHR can indicate whether the reaction will proceed to completion spontaneously or not.

As can be seen in Fig. 1, using our simple model the phenoxylation of Cl-AlPc does not seem to be thermodynamically favorable since the value of ΔHR is estimated to be approximately zero. The value for Br-AlPc is negative and for F-AlPc positive indicating that a phenoxylation reaction on Br-AlPc might be favorable whereas reaction on F-AlPc would not be favorable. It also means that in the presence of fluoride anions, the formation of F-AlPc from Cl-AlPc is more thermodynamically favorable than formation of the phenoxylated AlPc. This supports the observed formation of F-AlPc when pentafluoro phenol was used in DMF. Using this simple model, the reaction of X-SiPc and X-BsubPc (X = F, Cl, Br) have extremely negative ΔHR values and is therefore consistent with the ease by which each can be phenoxylated and further validates this simple model.


image file: c5ra04919a-f1.tif
Fig. 1 Heat of phenoxylation reaction (ΔHR) of different halide-MPcs.

While this basic model explains most of the observations when attempting to phenoxylate Cl-AlPc, the synthesis of F5PhO-AlPc and NO2PhO-AlPc did work. Given the ΔHR for phenoxylation of Cl-AlPc with each of these ‘acidic’ phenols is calculated to be near 0, subtle differences in the ΔHO–H and ΔHO–M values may cause a change in the sign of ΔHR and, hence, the spontaneity of the reaction. Subtle differences could be attributable to the acid dissociation constants (pKa); F5PhOH and NO2PhOH are 5.5 (ref. 60) and 7.15,62 respectively, compared to pKa's over 10 for alkylated phenols.62

Properties of phenoxy-AlPcs

Optical absorption measurements were taken for Cl-AlPc, NO2PhO-AlPc, and F5PhO-AlPc (Fig. 2). Minimal differences between the three compounds were observed. F5PhO-AlPc shows maximum absorption at λmax = 676 nm while Cl-AlPc and NO2PhO-AlPc have λmax = 680 and λmax = 678, respectively (DMSO solution). Solid-state absorbance spectra were taken from vacuum-deposited thin-films (∼1000 nm thick). The solid-state spectra were broader with peak absorptions slightly red-shifted from the solution spectra. Optical band gabs (EOG) were calculated using the onset of the main solid-state absorption peak. EOG ranged between 1.40–1.55 eV for the three AlPc compounds.
image file: c5ra04919a-f2.tif
Fig. 2 UV-Vis absorption spectra, solution in DMSO (solid line) and solid-state (dashed line), of Cl-AlPc, NO2PhO-AlPc, and F5PhO-AlPc.

Electrochemical oxidation and reduction potentials were measured using cyclic voltammetry (CV) and differential-pulse voltammetry (DPV) and are again consistent within this series of compounds (Fig. 3). Estimations of HOMO and LUMO levels were calculated using the empirical model developed by Dandrade et al.63 and the oxidation and reduction peaks from the DPV data (Table 1). Electrochemical energy gap (EEG) were defined and calculated as the difference between the oxidation and reduction peaks of the DPV data. The calculated values of NO2PhO-AlPc, and F5PhO-AlPc are similar to the energy levels of Cl-AlPc that are found in the literature.64


image file: c5ra04919a-f3.tif
Fig. 3 Cyclic voltammetry (blue) and differential-pulse voltammetry (red) of (a) F5PhO-AlPc and (b) NO2PhO-AlPc.

The thermal stability of all AlPc derivatives was measured using thermogravimetric analysis (TGA, Fig. 4). All of the three compounds showed two mass losses. Both phenoxylated AlPcs have a drop in mass at a lower temperature than the Cl-AlPc, between 350 °C and 400 °C compared to ∼500 °C for Cl-AlPc. The amount of mass lost corresponds approximately to the mass of the axial phenoxy groups and is therefore probably attributable to the detachment of the axial functionality. This analysis establishes the temperature window for device integration by vacuum deposition.


image file: c5ra04919a-f4.tif
Fig. 4 Thermogravimetric analysis (TGA) of Cl-AlPc, F5PhO-AlPc, and NO2PhO-AlPc.

Device integration

Multiple organic photovoltaic (OPV) devices were fabricated using Cl-AlPc, F5PhO-AlPc, and NO2PhO-AlPc as both electron donor and electron acceptor materials paired with either C60 or sexithiophene (α6T), respectively. C60 is a well-known electron acceptor material and α6-T is a well-known electron donor material. The application of each has been well documented in many OPV device structures.65–67 Cl-AlPc is also a relatively well studied organic semiconductor material that has been used as both an electron donor64,68 and an electron acceptor22 within OPVs, therefore is a good comparison point.

To begin, and as a point of comparison, we constructed baseline OPVs with active layer configurations of Cl-AlPc(20 nm)/C60(40 nm) and α6T (60 nm)/Cl-AlPc(20 nm). For Cl-AlPc(20 nm)/C60(40 nm) OPV the short-circuit current density (JSC) was measured at 3.85 mA cm−2, the open circuit voltage (VOC) at 0.67 V and the fill factor (FF) at 0.48 for an overall efficiency (η) of 1.23%. This is in line with the cell characteristics reported in the literature.64,68–71 For the α6T (60 nm)/Cl-AlPc(20 nm) cell the JSC was 3.82 mA cm−2, VOC was 0.61 V, FF 0.44 for an η of 1.02%. No point of comparison is available for the α6T/Cl-AlPc cell in the literature. The results are tabulated in Table 2 and the corresponding JV and EQE plots are illustrated in Fig. 5 and 6.

Table 2 The electrical performance and standard deviation of OPVs with Cl-AlPc, F5PhO-AlPc, and NO2PhO-AlPc as electron donors (paired with C60) and as electron acceptors (paired with α-6T)
Device Structure VOC (V) JSC (mA cm−2) FF η (%)
Cl-AlPc (20 nm)/C60 (40 nm) 0.67 ± 0.03 −3.85 ± 0.3 0.48 ± 0.03 1.23 ± 0.17
F5PhO-AlPc (20 nm)/C60 (40 nm) 0.8 ± 0.01 −2.97 ± 0.35 0.44 ± 0.01 1.06 ± 0.14
NO2PhO-AlPc(20 nm)C60 (40 nm) 0.82 ± 0.01 −2.63 ± 0.38 0.40 ± 0.01 0.86 ± 0.08
α-6T(60 nm)/Cl-AlPc(20 nm) 0.61 ± 0.009 −3.82 ± 0.5 0.44 ± 0.03 1.02 ± 0.07
α-6T(60 nm)/F5PhO-AlPc(20 nm) 0.67 ± 0.007 −4.2 ± 0.3 0.40 ± 0.02 1.12 ± 0.03
α-6T(60 nm)/NO2PhO-AlPc(20 nm) 0.66 ± 0.007 −2.75 ± 0.12 0.47 ± 0.02 0.84 ± 0.07



image file: c5ra04919a-f5.tif
Fig. 5 (a) The average JV curves and (b) the average external quantum efficiencies of Cl-AlPc/C60, F5PhO-AlPc/C60, and NO2PhO-AlPc/C60 OPVs.

image file: c5ra04919a-f6.tif
Fig. 6 (a) The average JV curves and (b) the average external quantum efficiencies of α-6T/Cl-AlPc, and α-6T/F5PhO-AlPc, and α-6T/NO2PhO- AlPc OPVs.

Moving to the phenoxylated derivatives as electron donating materials, for the F5PhO-AlPc(20 nm)/C60(40 nm) and NO2PhO-AlPc(20 nm)/C60(40 nm) OPVs, the JSC values were 2.92 and 2.63 mA cm−2, VOC values were 0.80 V and 0.82 V, FF values were 0.44 and 0.40 for overall η of 1.06% and 0.86%, respectively. When the results are compared against the baseline Cl-AlPc(20 nm)/C60(40 nm), the OPVs containing the phenoxylated derivatives had a VOC for the F5PhO-AlPc/C60 device 130 mV larger than for the Cl-AlPc/C60 device however, the JSC was 0.88 mA cm−2 smaller. The same general observations were seen for the NO2PhO-AlPc/C60 OPV with a VOC 132 mV larger than for Cl-AlPc/C60 and again a JSC 1.22 mA cm−2 lower than that of the Cl-AlPc/C60 OPV. The electrical characteristics illustrated in Fig. 5 and tabulated in Table 2.

Considering the application of F5PhO-AlPc, and NO2PhO-AlPc as electron accepting materials paired with α6T, the following OPVs were constructed: α6T(60 nm)/F5PhO-AlPc(20 nm) and α6-T(60 nm)/NO2PhO-AlPc(20 nm). The cells JSC values were 4.2 mA cm−2 and 2.75 mA cm−2, VOC values were 0.67 V and 0.66 V and FF values were 0.40 and 0.47 resulting in overall η of 1.12% and 0.84%, respectively (Table 2, Fig. 6). Notable, when F5PhO-AlPc and NO2PhO-AlPc used as electron acceptors and in line with their application as electron donating materials, an increase in VOC of 50–60 mV over α6T/Cl-AlPc was observed. The JSC of α6T/F5PhO-AlPc is 0.68 mA cm−2 larger than that of α6T/Cl-AlPc, however, the JSC of α6-T/NO2Pho-AlPc is 1.02 mA cm−2 smaller than that α6T/Cl-AlPc.

Based on the electrical characterization the following conclusions can be made regarding the application of phenoxylated AlPcs in OPVs. All of the AlPcs proved functional both as electron-donating and as electron-accepting materials in an OPV, therefore they are dually functional. There is an observed and consistent increase in VOC for the resulting OPVs when the phenoxylated AlPcs are used in place of Cl-AlPc and with one exception (α6T/F5PhO-AlPc) a decrease in the JSC.

These observations are in line with the ‘steric shielding effect’ described in both polymer72 and small molecule based OPVs.73,74 While this effect has been debated75 the concept is that addition of sterically bulky groups out of plane to the semiconducting π-electron system of an organic material enhances the VOC within an OPV. The VOC enhancement is attributed to the steric bulk increasing the distance between the electron donating and electron accepting materials at their interface thus resulting in a hole–electron pair having more potential. The high potential makes the separation into a hole–electron pair more difficult and therefore OPVs typically see a net drop in JSC.73

The results of the integration of the phenoxylated AlPc seem to add positively to the idea of the steric shielding effect. However, to positively state such a conclusion, further device optimization is being undertaken with an aim to improve the FF of the OPVs and to confirm that the reduced JSC (and overall efficiencies) cannot be overcome with more device engineering. Given our electrochemical data does indicate that there are energy differences in the HOMO and LUMO of Cl-AlPc, F5PhO-AlPc and NO2PhO-AlPc this could also be contributing to the differences seen in the VOC. Experiments will be undertaken in an effort to separate the relative contributions to the observed VOC differences.

Conclusions

We were successful in phenoxylating Cl-AlPc using pentafluoro phenol and p-nitrophenol to produce F5PhO-AlPc and NO2PhO-AlPc, respectively. A basic thermodynamic model was developed to explain limitations on conversion of the phenoxylation reaction when alkylated phenols are used. The optical, electrical, and thermal characteristics of the two novel AlPcs were determined. F5PhO-AlPc and NO2PhO-AlPcs were applied in OPVs both as electron donors and as electron acceptors paired with C60 and α6T respectively. When F5PhO-AlPc and NO2PhO-AlPc were applied as electron donating materials an increase in the VOC of the OPV of ∼130 mV was observed and ∼50–60 mV when applied as electron accepting materials. Given the increase in VOC these novel phenoxylated AlPcs show promise for application within OPVs although further device optimization and engineering is required to enhance the FF and JSC of the OPVs.

Acknowledgements

The authors would like to acknowledge financial support from Saudi Basic Industries (SABIC).

References

  1. N. B. McKeown, Phthalocyanine materials : synthesis, structure, and function, Cambridge University Press, Cambridge ; New York, 1998 Search PubMed.
  2. F. H. Moser and A. L. Thomas, Phthalocyanine compounds, Reinhold Pub. Co., New York, 1963 Search PubMed.
  3. X. Zhang, Y. Wang, Y. Ma, Y. Ye, Y. Wang and K. Wu, Langmuir, 2005, 22, 344–348 CrossRef PubMed.
  4. L. Cao, H. Chen, M. Wang, J. Sun, X. Zhang and F. Kong, J. Phys. Chem. B, 2002, 106, 8971–8975 CrossRef CAS.
  5. Y. Wang and D. Liang, Adv. Mater., 2010, 22, 1521–1525 CrossRef CAS PubMed.
  6. G. Williams, S. Sutty, R. Klenkler and H. Aziz, Sol. Energy Mater. Sol. Cells, 2014, 124, 217–226 CrossRef CAS PubMed.
  7. T. Jägeler-Hoheisel, F. Selzer, M. Riede and K. Leo, J. Phys. Chem. C, 2014, 118, 15128–15135 Search PubMed.
  8. W. Li, H. Yu, J. Zhang, Y. Yao, C. Wu and X. Hou, J. Phys. Chem. C, 2014, 118, 11928–11934 CAS.
  9. H. Fukui, S. Nakano, T. Uno, Q.-D. Dao, T. Saito, A. Fujii, Y. Shimizu and M. Ozaki, Org. Electron., 2014, 15, 1189–1196 CrossRef CAS PubMed.
  10. K. V. Chauhan, P. Sullivan, J. L. Yang and T. S. Jones, J. Mater. Chem., 2010, 6, 1173–1178 RSC.
  11. B. P. Rand, D. Cheyns, K. Vasseur, N. C. Giebink, S. Mothy, Y. Yi, V. Coropceanu, D. Beljonne, J. Cornil, J. L. Brédas and J. Genoe, Adv. Funct. Mater., 2012, 22, 2987–2995 CrossRef CAS PubMed.
  12. Z. Deng, Z. Lü, Y. Chen, Y. Yin, Y. Zou, J. Xiao and Y. Wang, Solid-State Electron., 2013, 89, 22–25 CrossRef CAS PubMed.
  13. K. Narayan, S. Varadharajaperumal, G. Mohan Rao, M. Manoj Varma and T. Srinivas, Curr. Appl. Phys., 2013, 13, 18–25 CrossRef PubMed.
  14. C. W. Tang, Appl. Phys. Lett., 1986, 48, 183–185 CrossRef CAS PubMed.
  15. N. M. Bamsey, A. P. Yuen, A.-M. Hor, R. Klenkler, J. S. Preston and R. O. Loutfy, Sol. Energy Mater. Sol. Cells, 2011, 95, 1970–1973 CrossRef CAS PubMed.
  16. N. Beaumont, I. Hancox, P. Sullivan, R. A. Hatton and T. S. Jones, Energy Environ. Sci., 2011, 5, 1708–1711 Search PubMed.
  17. D. K. Panda, F. S. Goodson, S. Ray and S. Saha, Chem. Commun., 2014, 50, 5358–5360 RSC.
  18. A. Hagfeldt, G. Boschloo, L. Sun, L. Kloo and H. Pettersson, Chem. Rev., 2010, 110, 6595–6663 CrossRef CAS PubMed.
  19. M.-E. Ragoussi, J.-H. Yum, A. K. Chandiran, M. Ince, G. de la Torre, M. Grätzel, M. K. Nazeeruddin and T. Torres, ChemPhysChem, 2014, 15, 1033–1036 CrossRef CAS PubMed.
  20. M.-E. Ragoussi, M. Ince and T. Torres, Eur. J. Org. Chem., 2013, 2013, 6475–6489 CrossRef CAS PubMed.
  21. J. Xue, B. P. Rand, S. Uchida and S. R. Forrest, Adv. Mater., 2005, 17, 66–71 CrossRef CAS PubMed.
  22. N. L. Beaumont, Doctor of Philosophy, University of Warwick, 2013.
  23. P. Sullivan, A. Duraud, L. Hancox, N. Beaumont, G. Mirri, J. H. R. Tucker, R. A. Hatton, M. Shipman and T. S. Jones, Adv. Energy Mater., 2011, 1, 352–355 CrossRef CAS PubMed.
  24. S. Schumann, R. A. Hatton and T. S. Jones, J. Phys. Chem. C, 2011, 115, 4916–4921 CAS.
  25. M. V. Martinez-Diaz, G. de la Torre and T. Torres, Chem. Commun., 2010, 46, 7090–7108 RSC.
  26. S. W. Cho, A. DeMasi, A. R. H. Preston, K. E. Smith, L. F. J. Piper, K. V. Chauhan and T. S. Jones, Appl. Phys. Lett., 2012, 100, 263302 CrossRef PubMed.
  27. I. Hancox, P. Sullivan, K. V. Chauhan, N. Beaumont, L. A. Rochford, R. A. Hatton and T. S. Jones, Org. Electron., 2010, 11, 2019–2025 CrossRef CAS PubMed.
  28. K. Vasseur, K. Broch, A. L. Ayzner, B. P. Rand, D. Cheyns, C. Frank, F. Schreiber, M. F. Toney, L. Froyen and P. Heremans, ACS Appl. Mater. Interfaces, 2013, 5, 8505–8515 CAS.
  29. E. R. L. Brisson, A. S. Paton, G. E. Morse and T. P. Bender, Ind. Eng. Chem. Res., 2011, 50, 10910–10917 CrossRef CAS.
  30. G. Löbbert, in Ullmann's Encyclopedia of Industrial Chemistry, Wiley-VCH Verlag GmbH & Co. KGaA, 2000,  DOI:10.1002/14356007.a20_213.
  31. S. M. O'Flaherty, S. V. Hold, M. J. Cook, T. Torres, Y. Chen, M. Hanack and W. J. Blau, Adv. Mater., 2003, 15, 19–32 CrossRef PubMed.
  32. B. A. Kamino and T. P. Bender, Dalton Trans., 2013, 42, 13145–13150 RSC.
  33. G. E. Morse, M. G. Helander, J. Stanwick, J. M. Sauks, A. S. Paton, Z.-H. Lu and T. P. Bender, J. Phys. Chem. C, 2011, 115, 11709–11718 CAS.
  34. C. A. Barker, K. S. Findlay, S. Bettington, A. S. Batsanov, I. F. Perepichka, M. R. Bryce and A. Beeby, Tetrahedron, 2006, 62, 9433–9439 CrossRef CAS PubMed.
  35. J. Guilleme, L. Martínez-Fernandez, D. Gonzalez-Rodríguez, I. Corral, M. Yáñez and T. Torres, J. Am. Chem. Soc., 2014, 136, 14289–14298 CrossRef CAS PubMed.
  36. M. V. Fulford, D. Jaidka, A. S. Paton, G. E. Morse, E. R. L. Brisson, A. J. Lough and T. P. Bender, J. Chem. Eng. Data, 2012, 57, 2756–2765 CrossRef CAS.
  37. G. E. Morse, J. S. Castrucci, M. G. Helander, Z.-H. Lu and T. P. Bender, ACS Appl. Mater. Interfaces, 2011, 3, 3538–3544 CAS.
  38. A. S. Paton, G. E. Morse, A. J. Lough and T. P. Bender, CrystEngComm, 2011, 13, 914–919 RSC.
  39. A. S. Paton, G. E. Morse, D. Castelino and T. P. Bender, J. Org. Chem., 2012, 77, 2531–2536 CrossRef CAS PubMed.
  40. J. Guilleme, D. González-Rodríguez and T. Torres, Angew. Chem., Int. Ed., 2011, 50, 3506–3509 CrossRef CAS PubMed.
  41. K. J. Wynne, Inorg. Chem., 1985, 24, 1339–1343 CrossRef CAS.
  42. K. J. Wynne, Inorg. Chem., 1984, 23, 4658–4663 CrossRef CAS.
  43. M. V. Fulford, A. J. Lough and T. P. Bender, Acta Crystallogr., Sect. B: Struct. Sci., 2012, 68, 636–645 CAS.
  44. Y. Yamasaki and K. Takaki, Dyes Pigm., 2006, 70, 105–109 CrossRef CAS PubMed.
  45. P. Samorí, H. Engelkamp, P. A. J. de Witte, A. E. Rowan, R. J. M. Nolte and J. P. Rabe, Adv. Mater., 2005, 17, 1265–1268 CrossRef PubMed.
  46. A. de la Escosura, M. V. Martínez-Díaz, T. Torres, R. H. Grubbs, D. M. Guldi, H. Neugebauer, C. Winder, M. Drees and N. S. Sariciftci, Chem.–Asian J., 2006, 1, 148–154 CrossRef CAS PubMed.
  47. N. B. McKeown, J. Mater. Chem., 2000, 10, 1979–1995 RSC.
  48. G. E. Morse, J. L. Gantz, K. X. Steirer, N. R. Armstrong and T. P. Bender, ACS Appl. Mater. Interfaces, 2014, 6, 1515–1524 CAS.
  49. B. H. Lessard, R. T. White, M. AL-Amar, T. Plint, J. S. Castrucci, D. S. Josey, Z.-H. Lu and T. P. Bender, ACS Appl. Mater. Interfaces, 2015, 7, 5076–5088 CAS.
  50. B. H. Lessard, M. AL-Amar, T. M. Grant, R. White, Z.-H. Lu and T. P. Bender, J. Mater. Chem. A, 2015, 3, 5047–5053 CAS.
  51. G. A. Kraus and S. J. V. Louw, Synlett, 1996, 1996, 726 CrossRef PubMed.
  52. N. Brasseur, R. Ouellet, C. L. Madeleine and J. E. V. Lier, Br. J. Cancer, 1999, 80, 1533–1541 CrossRef CAS PubMed.
  53. Y. Sakai, M. Ueda, A. Yahagi and N. Tanno, Polymer, 2002, 43, 3497–3503 CrossRef CAS.
  54. Bruker, APEX2, SAINT & SADABS, Bruker AXS Inc., Madison, Wisconsin, USA, 2007 Search PubMed.
  55. G. Sheldrick, Acta Crystallogr., Sect. A: Found. Crystallogr., 2008, 64, 112–122 CrossRef CAS PubMed.
  56. F. Santerre, R. Côté, G. Veilleux, R. G. Saint-Jacques and J. P. Dodelet, J. Phys. Chem., 1996, 100, 7632–7645 CrossRef CAS.
  57. H. Wagner, R. Loutfy and C.-K. Hsiao, J. Mater. Sci., 1982, 17, 2781–2791 CrossRef CAS.
  58. G. E. Morse, M. G. Helander, J. F. Maka, Z.-H. Lu and T. P. Bender, ACS Appl. Mater. Interfaces, 2010, 2, 1934–1944 CAS.
  59. J. E. Owen and M. E. Kenney, Inorg. Chem., 1962, 1, 331–333 CrossRef CAS.
  60. L. C. T. Shoute, J. P. Mittal and P. Neta, J. Phys. Chem., 1996, 100, 3016–3019 CrossRef CAS.
  61. Y.-R. Luo and J.-P. Cheng, Bond Dissociation Energies, Handbook of Chemistry and Physics, edn, 2012, pp. 9–65 Search PubMed.
  62. Dissociation Constants of Organic Acids and Bases, Handbook of Chemistry and Physics, edn, 2014, pp. 5–94 Search PubMed.
  63. B. Dandrade, S. Datta, S. Forrest, P. Djurovich, E. Polikarpov and M. Thompson, Org. Electron., 2005, 6, 11–20 CrossRef CAS PubMed.
  64. D. Y. Kim, F. So and Y. Gao, Sol. Energy Mater. Sol. Cells, 2009, 93, 1688–1691 CrossRef CAS PubMed.
  65. C.-Z. Li, H.-L. Yip and A. K. Y. Jen, J. Mater. Chem., 2012, 22, 4161–4177 RSC.
  66. Y. He and Y. Li, Phys. Chem. Chem. Phys., 2011, 13, 1970–1983 RSC.
  67. K. Cnops, B. P. Rand, D. Cheyns, B. Verreet, M. A. Empl and P. Heremans, Nat. Commun., 2014, 5, 3406 Search PubMed.
  68. R. F. Bailey-Salzman, B. P. Rand and S. R. Forrest, Appl. Phys. Lett., 2007, 91, 013508 CrossRef PubMed.
  69. V. Chauhan, R. Hatton, P. Sullivan, T. Jones, S. W. Cho, L. Piper, A. deMasi and K. Smith, J. Mater. Chem., 2010, 20, 1173–1178 RSC.
  70. N. Beaumont, S. W. Cho, P. Sullivan, D. Newby, K. E. Smith and T. S. Jones, Adv. Funct. Mater., 2012, 22, 561–566 CrossRef CAS PubMed.
  71. I. Hancox, K. V. Chauhan, P. Sullivan, R. A. Hatton, A. Moshar, C. P. A. Mulcahy and T. S. Jones, Energy Environ. Sci., 2010, 3, 107–110 CAS.
  72. L. Yang, H. Zhou and W. You, J. Phys. Chem. C, 2010, 114, 16793–16800 CAS.
  73. P. Erwin and M. E. Thompson, Appl. Phys. Lett., 2011, 98, 223305 CrossRef PubMed.
  74. S. L. Lam, Z. Liu, F. Zhao, C.-L. K. Less and W. L. Kwan, Chem. Commun., 2013, 49, 4543 RSC.
  75. K. R. Graham, P. Ewan, D. Nordland, K. Vandewal, R. Li, G. O. Ngongang, E. T. Hoke, A. Salleo, M. E. Thompson, M. D. McGehee and A. Amassian, Adv. Mater., 2013, 25, 6076–6082 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available: High and low resolution mass spectra and crystallographic data. CCDC 1052023 and 1061523. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c5ra04919a

This journal is © The Royal Society of Chemistry 2015