Next Article in Journal
Preparation of Modified Beeswax and Its Influence on the Surface Properties of Compressed Poplar Wood
Previous Article in Journal
Review on Microwave-Matter Interaction Fundamentals and Efficient Microwave-Associated Heating Strategies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermally Stable Solution Processed Vanadium Oxide as a Hole Extraction Layer in Organic Solar Cells

1
Department of Physics & Astronomy, University of Sheffield, Hicks Building, Hounsfield Rd., Sheffield, South Yorkshire S3 7RH, UK
2
Department of Chemistry, University of Sheffield, Sheffield, South Yorkshire S3 7HF, UK
*
Author to whom correspondence should be addressed.
Materials 2016, 9(4), 235; https://doi.org/10.3390/ma9040235
Submission received: 16 February 2016 / Revised: 12 March 2016 / Accepted: 22 March 2016 / Published: 25 March 2016
(This article belongs to the Section Energy Materials)

Abstract

:
Low-temperature solution-processable vanadium oxide (V2Ox) thin films have been employed as hole extraction layers (HELs) in polymer bulk heterojunction solar cells. V2Ox films were fabricated in air by spin-coating vanadium(V) oxytriisopropoxide (s-V2Ox) at room temperature without the need for further thermal annealing. The deposited vanadium(V) oxytriisopropoxide film undergoes hydrolysis in air, converting to V2Ox with optical and electronic properties comparable to vacuum-deposited V2O5. When s-V2Ox thin films were annealed in air at temperatures of 100 °C and 200 °C, OPV devices showed similar results with good thermal stability and better light transparency. Annealing at 300 °C and 400 °C resulted in a power conversion efficiency (PCE) of 5% with a decrement approximately 15% lower than that of unannealed films; this is due to the relative decrease in the shunt resistance (Rsh) and an increase in the series resistance (Rs) related to changes in the oxidation state of vanadium.

1. Introduction

Currently, the high cost of commercial inorganic photovoltaics remains an obstacle for the wide-scale installation in both residential and commercial settings [1]. In order to overcome this high cost, researchers have studied numerous materials as alternatives, with organic polymer solar cells being a promising candidate [2,3]. This is due to organic polymers exhibiting several advantageous properties, such as solution-processability, mechanical-flexibility, and thin device architectures allowing for light-weight devices [4,5,6]. This combination of factors allows organic solar to scale up manufacture via roll-to-roll- or sheet-to-sheet-based deposition techniques leading to reduced fabrication costs. Recent advances in polymer synthesis and device processing have pushed the power conversion efficiency of single junction cells as high as 10.8% [7]. These efficiencies are above the 10% mark often quoted as the point at which polymer photovoltaics will become commercially viable. However, currently, the lifetime of devices incorporating the highest performing materials are below the 10 year mark required; one method to improve the lifetime of devices is to replace the commonly-used hole extraction interfacial layer (HELs), polyethylene dioxythiophene:polystyrenesulfonate (PEDOT:PSS) [8,9,10]. This is because, despite the positive aspects of aqueous PEDOT:PSS, the residual moisture and the acidic nature of PEDOT:PSS can cause degradation of the electrode and organic films and, therefore, reduce the devices operational lifetime [11,12,13,14]. To overcome some of these issues thin metal oxides such as MoO3, WO3, NiO, and V2O5 have been used; these materials have been shown to exhibit performances that are similar to or better than OPVs with an interfacial PEDOT:PSS layer [15,16,17,18]. Not only are these materials of interest in organic photovoltaics because of their unique electronic properties and chemical stability, they are also being used in other photovoltaic technologies such as CiGS, and CdTe [19,20,21,22]. However, like PEDOT:PSS the presence of water can lead to changes at the interface between metal oxides and organic semiconductors, due to the intercalation of water within metal oxide clusters leading to changes within the energy level structure [23]. This makes the improvement of processing conditions and reduction in the cost and complexity of depositing these interfacial layers highly desirable.
Solution-processing of the metal oxide interfacial layers has shown significant promise for reducing the complexity of the deposition of these materials. Previous studies have shown that vanadium oxide can be deposited from solution at low temperature in air without the need for any high temperature post-deposition treatments [24,25,26,27,28]. On the other hand, some optoelectronic devices which are fabricated at high temperatures require developed metal oxides which must be able to keep their properties under high temperature fabrication processes. Whilst the optical and chemical properties of V2Ox have previously been investigated after high-temperature annealing [29,30,31], its performance with OPVs has not. Therefore, our motivation for this work is to explore the impact of thermal heating on V2Ox thin films as HELs in organic solar cells.
In this paper we incorporated V2Ox thin films deposited from a vanadium(V) oxytriisopropoxide precursor into PFDT2BT-8:PC70BM OPV devices achieving a power conversion efficiency of 6%, comparable to PEDOT:PSS and vacuum deposited MoO3. Furthermore, we fabricated some OPV devices with annealed V2Ox layers at high temperatures in air before spin coating the active layer to study their thermal stability. We demonstrated that OPV devices show PCE ≥ 5% with thermally-annealed V2Ox at 400 °C. Absorption spectroscopy, X-ray photoelectron spectroscopy (XPS), and ultraviolet photoelectron spectroscopy (UPS) are combined to explain the J–V characteristics of OPVs.

2. Results and Discussion

2.1. PFDT2BT-8 Structure

The chemical structure of our donor polymer used in this work is shown in Figure 1a [32]. The highest occupied molecular orbital (HOMO) energy level and lowest unoccupied molecular orbital (LUMO) of PFDT2BT-8 are −5.33 eV and −3.34 eV, respectively, as determined from a cyclic voltammetry. The energy band diagram and work function of the relative materials in our study are presented in Figure 1b [32,33].

2.2. s-V2Ox as a Hole Extraction Layer

To assess the performance of PFDT2BT-8 polymer with the s-V2Ox interlayer, we fabricated sets of PFDT2BT-8:PC70BM devices with variable thicknesses of s-V2Ox. The extracted data suggests an optimum s-V2Ox layer thickness below 10 nm. Therefore, we selected a layer thickness of 5 nm for all s-V2Ox devices.
Figure 2 shows the J–V characteristics of our OPV devices and those of the most widely used HELs, PEDOT:PSS and thermally evaporated MoO3 under AM 1.5 G illumination. It can be seen that using PEDOT:PSS and s-V2Ox layers as HEL showed a similar photovoltaic response (PCE = 6.5%) which is better than devices fabricated with MoO3 interlayer (PCE = 6.3%). These results indicate that a high performance can be achieved for PFDT2BT-8 devices by using untreated s-V2Ox films which are in good agreement with the previous studies in literature [25,27,34].
To explore the effect of thermal annealing (≥100 °C) on the photovoltaic response of our devices, s-V2Ox films were annealed in air at temperatures of 100, 200, 300, and 400 °C for 30 min before spin coating the active layer. The corresponding current-voltage characteristic of OPVs is shown in Figure 3. The photovoltaic parameters obtained are summarised in Table 1 which represent the average of at least 12 pixels from 18 pixels defined on three separate substrates. The errors quoted are defined by the standard deviation about the mean.
It can be seen that the performance of devices annealed at temperatures 100 or 200 °C is quite similar to those prepared without heat treatment. The slight increase in PCE of devices annealed up to 200 °C can be ascribed to the relative increase in the shunt resistance Rsh and decrease in the series resistance Rs. In contrast, as the film annealing temperature is raised to 400 °C, OPV devices show a PCE of 5% with a decrement approximately 15% lower than that of those with unannealed films. It is clear from Table 1 that the relatively weak photovoltaic response can be attributed to the decrease in open-circuit voltage, due to the change in chemical structure as will be discussed based on the XPS data later on. Furthermore, the shunt resistance falls considerably on annealing above 300 °C reaching a minimum Rsh of 933 Ω·cm2 on annealing at 400 °C due to the recombination process at the anode interface. This caused a decrease in hole density at the anode interface leading to a decrease in the internal electric field [35,36].

2.3. Optical Properties

To understand the changes that we observed in the photovoltaic response as a function of the annealing temperature, we investigated the optical properties of annealed s-V2Ox films. Figure 4 presents the optical transmittance of the films as a function of the wavelength for the as-deposited and annealed films. All s-V2Ox films show high transmission for wavelengths above 500 nm covering the majority of the solar spectrum. The transmittance in the visible wavelength range, from 400 to 500 nm, increases slightly by annealing up to 300 °C which is attributed to change of the films refractive index related to chemical structure changes. On annealing to 400 °C, the film exhibits absorption peak at 415 nm which could be ascribed to small polaron absorption [37,38]. This effect results from disordering defects in V2Ox structure leading to transferring charges between neighbouring sites with a significant spread in energy.
Using Tauc’s Law, the optical band gap of s-V2Ox films was determined [25,27]. The relationship between the band gap energy ( E g ) and the absorption coefficient (α) can be represented as:
α ( h ν ) [ h ν   E g ] 1 N
where h ν represents the photon energy and N is equal to 2. As can be seen in Figure 4, the optical absorption coefficient ( α h ν ) 2   is plotted as a function of the incident photon energy. By extrapolating from the straight-line portion of the plots to the energy-axis, we determined that E g of the unannealed film lies at 2.5 eV which is very close to those reported in literature for solution-processed samples [24,27,39]. Several studies have reported various values of V2Ox band gap due to the different processing conditions or the measurement method. For instance, electron spectroscopy measurements of V2Ox layer spin coated in air from isopropanol solution of vanadium(V) oxitriisopropoxide showed a band gap of 3.6 eV [24]. In contrast, the optical absorption experiments of the same precursor revealed a lower band gap of 2.3 eV [25]. In addition, impact of the processing conditions on the V2Ox band gap has been reported by other authors [24,27,39,40,41]. For example, V2Ox samples spin coated and annealed under N2 atmosphere had a band gap of 3.2 eV while fabrication of the films in air showed lower band gaps of 2.5 eV [40]. On annealing s-V2Ox films to 200 °C, the band gap energy rises to 2.78 eV. On further annealing, E g reduces again to 2.42 eV at an annealing temperature of 400 °C. Looking at the individual values of Jsc in Table 1 it can be observed that the variation in the band gap is independent of the OPV performance.

2.4. Atomic Force Microscopy

AFM scanning was performed to investigate the impact of thermal annealing on surface topography of s-V2Ox. Figure 5 shows 2 µm × 2 µm AFM scans for an unannealed film (a), and films thermally-annealed at 200 °C and 400 °C for 30 min ((b) and (c), respectively). RMS roughness of films decreased slightly upon annealing, from 2.5 nm for an as-cast precursor film to 1.9 nm for films annealed at 400 °C, with no local crystallisation structure observed. The low RMS of V2Ox annealed samples avoided short circuit faults when they were incorporated into OPV devices as shown in Table 1.

2.5. X-ray Photoelectron Spectroscopy

We investigated the chemical composition and electronic structure of s-V2Ox thin films by photoelectron spectroscopy (UPS and XPS). Figure 6a shows the XPS spectrum of an as-cast precursor film prepared in ambient conditions. The O1s signal, determined at 530.0 eV as reported in literature [42], was used as a binding energy reference. Therefore, we find V2p1/2, and V2p3/2 lines correspond to 517.1 eV and 524.6 eV, respectively, which is in good agreement with the commonly reported values [42,43,44,45,46].
Decomposition analysis reveals that V2p3/2 peak consists of two different species (i.e., V5+ and V4+ oxidation state) with a ratio of 5.7:1. The peak position Of V4+ and V5+ in V2p3/2 spectrum was identified by using L-G fitting to be 515.8 eV and 517.1 eV, respectively, in which V4+ oxidation state arises as a low-energy shoulder in the core level of V2p3/2. The calculated composition analyses indicate that a small amount of oxygen vacancy exists in unannealed s-V2Ox films. Fabrication conditions such as atmospheric gases and thermal annealing can increase the concentration of V4+ species leading to the reduction of the V5+ to V4+. For instance, H2O molecules in air facilitate the oxygen removal such that oxygen-hydrogen interaction weakens the binding energy of the oxygen atom with the neighbouring vanadium atoms [47,48]. Moreover, a previous study demonstrated that V2O5 can be reduced during XPS measuring in which X-rays create vanadyl ( V = O ) oxygen atoms vacancies [49].
Figure 6b exhibits a comparison between the XPS spectra of (i) unannealed s-V2Ox films and those annealed in air at three different temperatures (ii) 200 °C, (iii) 300 °C, and (iv) 400 °C. As observed, the binding energies of the V2p3/2 and O1s core levels do not change appreciably with increasing temperature. Furthermore, as the annealing temperature increases, the V 2p3/2 spectrum becomes broader at low binding energies due to the increase of V4+ species. Table 1 summarises the change of the oxidation state of vanadium with annealing temperature which is consistent with previous studies [50,51]. By deconvolution of V2p3/2 lines, we did not observe any shift in the binding energy of V4+ and V5+ peaks in any cases. In addition, it is evident that V4+ oxidation state rises steadily with increasing annealing temperatures. Therefore, V2Ox thin films are gradually reduced owing to the generation of more oxygen vacancies. This partial reduction leads to an increase in the V3d electron density [52]. Furthermore, oxygen vacancies in V2Ox can cause localisation excess electrons in unfilled 3d orbitals which can explain the absorption peak near to the ultraviolet band observed in Figure 4 [37]. From Table 2 we conclude that vanadium oxide can be partially reduced by a thermal annealing process in air for 30 min from V2O4.93 to V6O14. Increasing the annealing temperature can result in further reduction as reported previously [50].
In the literature, there have been detailed investigations on the physical properties and gradual reduction of vanadium oxide by heating at high temperatures [29,50].

2.6. Ultraviolet Photoelectron Spectroscopy

Figure 7 plots the UPS spectra of V2Ox films that have been annealed at different temperatures. The work function and valence band edge determined by the UPS are also summarised in Table 1. As shown in Figure 7a, the position of the secondary electron cut-off for unannealed film was determined to be at the binding energy of 15.94 eV below the Fermi level, corresponding to a work function of 5.26 eV. Compared to previous studies, the work function of our s-V2Ox films is similar to those prepared by different processes [34]. On annealing, the secondary electron cut-off shows a slight shift up to 0.07 eV toward the higher binding energy. While we see a 0.07 eV change in the work function we can’t assign these changes directly to the processing of the s-V2Ox due to the possible influence of adsorbates. Figure 7b displays the onset of the s-V2Ox valence band to be 2.5 eV below the Fermi level which is identical with all samples. We did not notice any shift for the valence band maximum with increase annealing temperature. Careful observation of the valence band region of films annealed at 300 °C and 400 °C reveals the formation of more gap states at 1 eV below the Fermi level as shown in the inset of Figure 7b. This suggests that oxygen vacancies shift the Fermi level up toward the conduction band leaving some occupied states within the band gap. Therefore, introducing oxygen vacancies can act as n-type dopants resulting in a decrease in the work function of vanadium oxide [53]. Consequently, photoelectron spectroscopy analysis confirms that the high-temperature annealing of s-V2Ox film produces a change in the chemical structure of the vanadium oxide layer and, thus, increases slightly the hole-electron recombination at the interface. Nevertheless, our experiments demonstrate that s-V2Ox film can be used to replace evaporated metal oxides in optoelectronic devices which are fabricated at high temperatures.

3. Materials and Methods

3.1. Materials

Vanadium(V) oxytriisopropoxide was purchased from Sigma-Aldrich (Dorset, UK) and was mixed with iso-propanol (99.5%) at a 1:250 volume ratio to obtain a solution with a concentration of 4 mg·mL−1. PEDOT:PSS was purchased from Ossila Ltd., Sheffield, UK, MoO3 (99.95%) was purchased from Testbourne Ltd (Basingstoke, UK), aluminium (99.99%) and calcium (99%) were purchased from Sigma-Aldrich (Dorset, UK). The donor polymer PFDT2BT-8 was synthesised in the Department of Chemistry at the University of Sheffield via a previously reported method, and had a molecular weight of 91.6 kDa and a polydispersity index (PDI) of 1.47 [32]. PC70BM was purchased from Ossila Ltd with a purity of 95% (5% PC60BM). The active layer solution was prepared by mixing PFDT2BT-8 and PC70BM at a weight ratio of 1:4 in chloroform with an overall concentration of 20 mg·mL−1.

3.2. OPV Device Fabrication

All OPV devices were fabricated onto pre-patterned ITO coated glass substrates purchased from Ossila Ltd (Sheffield, UK). Prior to use, the substrates were sonicated in a warm cleaning solution of Hellmanex (2 wt %) for 10 min at 70 °C. After sonication, they were washed with the de-ionized water. They were then placed in isopropanol and sonicated for 10 min at 70 °C. Finally, they were dried with nitrogen gas. Thin films (~5 nm) of s-V2Ox were deposited via spin coating onto cleaned ITO substrates in ambient conditions and were then transferred into a dry nitrogen glove box. A thermal evaporated MoO3 film (10 nm) was deposited at a rate of 3 Å·s−1 and at a pressure of ~10−7 mbar in a high-vacuum system in the glove box. PEDOT:PSS film with (30 nm ± 3 nm) thick was spin cast in ambient conditions and was then annealed at 130 °C for an hour in the glove box environment. Semiconducting thin films were prepared by spin casting the solution onto a substrate at a spin speed of 3000 rpm in order to obtain an active layer with a thickness of 70 nm ± 4 nm. The substrates were then transferred within the glove box to a high-vacuum system (10−7 mbar) to thermally evaporate the cathodes. The bi-layer cathodes of Ca (5 nm) and Al (100 nm) were evaporated at a rate of 3 Å·s−1 and 10 Å·s−1, respectively. The final step in device fabrication was encapsulation of the central area of each substrate by using a glass slide and light-curable epoxy, to extend the lifetime for measurement and storage.

3.3. Current-Density Characterisation

OPV devices were measured under ambient conditions using a Keithley 2400 source meter (Tektronix Ltd., Bracknell, UK) and a Newport 92251A-1000 AM1.5 solar simulator (Newport Co.,LTD, Didcot, UK). An NREL calibrated silicon diode was used to calibrate the power output at 1000 W·cm−2. Each device had an area of approximately 2.6 mm2 as defined by a shadow mask.

3.4. Atomic Force Microscopy

A Veeco Instruments Dimension 3100 was used to perform the AFM. Aluminium-coated silicon tips from Budget Sensors (Tap 300 Al-G) with a resonance frequency of 300 kHz and a spring constant of 40 N·m−1 were used throughout. The obtained data were processed using the Gwyddion.

3.5. Photoelectron Spectroscopy

The UPS and XPS measurements were carried out using KratosUltra AXIS photoelectron spectroscopy (Kratos Analytical Ltd., Manchester, UK). XPS measurements were taken using the Al Kα radiation with an energy of 1486.6 eV as the excitation source, a band pass energy of 10 eV, a step size of 0.025 eV, and a dwell time of 250 ms. UPS measurements were taken using the He (I) emission line (21.2 eV), a band pass energy of 10 eV, a step size of 0.025 eV, and a dwell time of 250 ms.

3.6. Absorption Spectroscopy

Absorption measurements were conducted under ambient conditions using a Horiba Fluoromax 4 spectrometer (HORIBA Ltd., Stanmore Middlesex, UK) with a xenon arc lamp (200 to 950 nm). Transmittance and absorbance of all the samples in this work were measured with the range 300–900 nm in increments of 2 nm and a slit size of 2 nm. Transmittance spectra were normalised against the power output recorded by the reference detector to account for any potential variations in light output between the separate measurements.

4. Conclusions

In contrast to previous studies, we have demonstrated that thermal annealing of V2Ox thin films deposited from a vanadium(V) oxytriisopropoxide does not significantly enhance the performance of OPV devices. However, the efficiency of OPV devices is reduced by 15% after annealing s-V2Ox layers at high temperatures (i.e., 400 °C) due to the decrease of Voc and shunt resistance. Absorption spectroscopy studies reveal that OPV performance is independent of the variable optical band gap of s-V2Ox. XPS analysis shows that annealing of vanadium oxide films for 30 min in air resulted in a reduction of V2O4.93 to V6O14 as a result of the generation of more oxygen vacancies. This reduction causes a decrease of Voc and, thus, the PCE of OPV device. Nevertheless, OPV results confirm that s-V2Ox thin film is suitable for optoelectronic devices which are fabricated at high temperatures, and is promising for efficient OPV with a low cost manufacturing process.

Acknowledgments

We thank the UK EPSRC for funding the project “Polymer fullerene photovoltaic devices: New materials and innovative processes for high-volume manufacture” (EP/I02864/1) and “Photovoltaics for Future Societies” (EP/I032541/1). A. Alsulami thanks the Saudi Cultural Bureau in London, UK, and King Abdulaziz City for Science and Technology (KACST) in Riyadh, Saudi Arabia, for the provision of a PhD scholarship.

Author Contributions

Abdullah Alsulami, Jonathan Griffin and Rania Alqurashi fabricated the OPV devices, performed the experiments and contributed to manuscript writing. Ahmed Iraqi and Hunan Yi synthesised PFDT2BT-8, measured its electronic structure and commented on various aspects during the manuscript preparation. Alastair Buckley and David Lidzey conceived and designed the overall research concept, analyzed the data, contributed to manuscript writing and supervised the work.

Conflicts of Interest

The authors declare no conflict of interest. The founding sponsors had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, and in the decision to publish the results.

References

  1. Timilsina, G.R.; Kurdgelashvili, L.; Narbel, P.A. Solar energy: Markets, economics and policies. Renew. Sustain. Energy Rev. 2012, 16, 449–465. [Google Scholar] [CrossRef]
  2. Nielsen, T.D.; Cruickshank, C.; Foged, S.; Thorsen, J.; Krebs, F.C. Business, market and intellectual property analysis of polymer solar cells. Sol. Energy Mater. Sol. Cells 2010, 94, 1553–1571. [Google Scholar] [CrossRef]
  3. Li, G.; Zhu, R.; Yang, Y. Polymer solar cells. Nat. Photonics 2012, 6, 153–161. [Google Scholar] [CrossRef]
  4. Shah, A.; Torres, P.; Tscharner, R.; Wyrsch, N.; Keppner, H. Photovoltaic technology: The case for thin-film solar cells. Science 1999, 285, 692–698. [Google Scholar] [CrossRef] [PubMed]
  5. Guenes, S.; Neugebauer, H.; Sariciftci, N.S. Conjugated polymer-based organic solar cells. Chem. Rev. 2007, 107, 1324–1338. [Google Scholar] [CrossRef] [PubMed]
  6. Delgado, J.L.; Bouit, P.A.; Filippone, S.; Herranz, M.A.; Martin, N. Organic photovoltaics: A chemical approach. Chem. Commun. 2010, 46, 4853–4865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Liu, Y.; Zhao, J.; Li, Z.; Mu, C.; Ma, W.; Hu, H.; Jiang, K.; Lin, H.; Ade, H.; Yan, H. Aggregation and morphology control enables multiple cases of high-efficiency polymer solar cells. Nat. Commun. 2014, 5. [Google Scholar] [CrossRef] [PubMed]
  8. Steim, R.; Kogler, F.R.; Brabec, C.J. Interface materials for organic solar cells. J. Mater. Chem. 2010, 20, 2499–2512. [Google Scholar] [CrossRef]
  9. Arias, A.C.; Granstrom, M.; Thomas, D.S.; Petritsch, K.; Friend, R.H. Doped conducting-polymer-semiconducting-polymer interfaces: Their use in organic photovoltaic devices. Phys. Rev. B 1999, 60, 1854–1860. [Google Scholar] [CrossRef]
  10. Ko, C.J.; Lin, Y.K.; Chen, F.C.; Chu, C.W. Modified buffer layers for polymer photovoltaic devices. Appl. Phys. Lett. 2007, 90. [Google Scholar] [CrossRef]
  11. Lee, K.; Kim, J.Y.; Park, S.H.; Kim, S.H.; Cho, S.; Heeger, A.J. Air-stable polymer electronic devices. Adv. Mater. 2007, 19. [Google Scholar] [CrossRef]
  12. Norrman, K.; Madsen, M.V.; Gevorgyan, S.A.; Krebs, F.C. Degradation patterns in water and oxygen of an inverted polymer solar cell. J. Am. Chem. Soc. 2010, 132, 16883–16892. [Google Scholar] [CrossRef] [PubMed]
  13. Ratcliff, E.L.; Zacher, B.; Armstrong, N.R. Selective inter layers and contacts in organic photovoltaic cells. J. Phys. Chem. Lett. 2011, 2, 1337–1350. [Google Scholar] [CrossRef] [PubMed]
  14. De Jong, M.P.; van Ijzendoorn, L.J.; de Voigt, M.J.A. Stability of the interface between indium-tin-oxide and poly(3,4-ethylenedioxythiophene)/poly(styrenesulfonate) in polymer light-emitting diodes. Appl. Phys. Lett. 2000, 77, 2255–2257. [Google Scholar] [CrossRef]
  15. Shrotriya, V.; Li, G.; Yao, Y.; Chu, C.W.; Yang, Y. Transition metal oxides as the buffer layer for polymer photovoltaic cells. Appl. Phys. Lett. 2006, 88. [Google Scholar] [CrossRef]
  16. Park, J.H.; Lee, T.W.; Chin, B.D.; Wang, D.H.; Park, O.O. Roles of interlayers in efficient organic photovoltaic devices. Macromol. Rapid Commun. 2010, 31, 2095–2108. [Google Scholar] [CrossRef] [PubMed]
  17. Chen, S.; Manders, J.R.; Tsang, S.W.; So, F. Metal oxides for interface engineering in polymer solar cells. J. Mater. Chem. 2012, 22, 24202–24212. [Google Scholar] [CrossRef]
  18. Wang, H.Q.; Li, N.; Guldal, N.S.; Brabec, C.J. Nanocrystal V2O5 thin film as hole-extraction layer in normal architecture organic solar cells. Org. Electron. 2012, 13, 3014–3021. [Google Scholar] [CrossRef]
  19. Chambers, B.A.; MacDonald, B.I.; Ionescu, M.; Deslandes, A.; Quinton, J.S.; Jasieniak, J.J.; Andersson, G.G. Examining the role of ultra-thin atomic layer deposited metal oxide barrier layers on cdte/ito interface stability during the fabrication of solution processed nanocrystalline solar cells. Sol. Energy Mater. Sol. Cells 2014, 125, 164–169. [Google Scholar] [CrossRef]
  20. Townsend, T.K.; Yoon, W.; Foos, E.E.; Tischler, J.G. Impact of nanocrystal spray deposition on inorganic solar cells. ACS Appl. Mater. Interfaces 2014, 6, 7902–7909. [Google Scholar] [CrossRef] [PubMed]
  21. Kim, A.; Won, Y.; Woo, K.; Jeong, S.; Moon, J. All-solution-processed indium-free transparent composite electrodes based on Ag nanowire and metal oxide for thin- film solar cells. Adv. Funct. Mater. 2014, 24, 2462–2471. [Google Scholar] [CrossRef]
  22. Song, S.H.; Aydil, E.S.; Campbell, S.A. Metal-oxide broken-gap tunnel junction for copper indium gallium diselenide tandem solar cells. Sol. Energy Mater. Sol. Cells 2015, 133, 133–142. [Google Scholar] [CrossRef]
  23. Gwinner, M.C.; di Pietro, R.; Vaynzof, Y.; Greenberg, K.J.; Ho, P.K.H.; Friend, R.H.; Sirringhaus, H. Doping of organic semiconductors using molybdenum trioxide: A quantitative time-dependent electrical and spectroscopic study. Adv. Funct. Mater. 2011, 21, 1432–1441. [Google Scholar] [CrossRef]
  24. Zilberberg, K.; Trost, S.; Meyer, J.; Kahn, A.; Behrendt, A.; Luetzenkirchen-Hecht, D.; Frahm, R.; Riedl, T. Inverted organic solar cells with sol-gel processed high work-function vanadium oxide hole-extraction layers. Adv. Funct. Mater. 2011, 21, 4776–4783. [Google Scholar] [CrossRef]
  25. Zilberberg, K.; Trost, S.; Schmidt, H.; Riedl, T. Solution processed vanadium pentoxide as charge extraction layer for organic solar cells. Adv. Energy Mater. 2011, 1, 377–381. [Google Scholar] [CrossRef]
  26. Tan, Z.A.; Zhang, W.Q.; Cui, C.H.; Ding, Y.Q.; Qian, D.P.; Xu, Q.; Li, L.J.; Li, S.S.; Li, Y.F. Solution-processed vanadium oxide as a hole collection layer on an ito electrode for high-performance polymer solar cells. Phys. Chem. Chem. Phys. 2012, 14, 14589–14595. [Google Scholar] [CrossRef] [PubMed]
  27. Xie, F.X.; Choy, W.C.H.; Wang, C.D.; Li, X.C.; Zhang, S.Q.; Hou, J.H. Low-temperature solution-processed hydrogen molybdenum and vanadium bronzes for an efficient hole-transport layer in organic electronics. Adv. Mater. 2013, 25, 2051–2055. [Google Scholar] [CrossRef] [PubMed]
  28. Jin, S.; Jung, B.J.; Song, C.K.; Kwak, J. Room-temperature and solution-processed vanadium oxide buffer layer for efficient charge injection in bottom-contact organic field-effect transistors. Curr. Appl. Phys. 2014, 14, 1809–1812. [Google Scholar] [CrossRef]
  29. Oksuzoglu, R.M.; Bilgic, P.; Yildirim, M.; Deniz, O. Influence of post-annealing on electrical, structural and optical properties of vanadium oxide thin films. Opt. Laser Technol. 2013, 48, 102–109. [Google Scholar] [CrossRef]
  30. Sahana, M.B.; Sudakar, C.; Thapa, C.; Lawes, G.; Naik, V.M.; Baird, R.J.; Auner, G.W.; Naik, R.; Padmanabhan, K.R. Electrochemical propertiesof V2O5 thin films deposited by spin coating. Mater. Sci. Eng. B 2007, 143, 42–50. [Google Scholar] [CrossRef]
  31. Haber, J.; Witko, M.; Tokarz, R. Vanadium pentoxide.1. Structures and properties. Appl. Catal. A Gen. 1997, 157, 3–22. [Google Scholar] [CrossRef]
  32. Watters, D.C.; Yi, H.; Pearson, A.J.; Kingsley, J.; Iraqi, A.; Lidzey, D. Fluorene-based co-polymer with high hole mobility and device performance in bulk heterojunction organic solar cells. Macromol. Rapid Commun. 2013, 34, 1157–1162. [Google Scholar] [CrossRef] [PubMed]
  33. Lu, L.; Xu, T.; Chen, W.; Landry, E.S.; Yui, L. Ternary blend polymer solar cells with enhanced power conversion efficiency. Nat. Photonics 2014, 8, 716–722. [Google Scholar] [CrossRef]
  34. Zilberberg, K.; Meyer, J.; Riedl, T. Solution processed metal-oxides for organic electronic devices. J. Mater. Chem. C 2013, 1, 4796–4815. [Google Scholar] [CrossRef]
  35. Wagenpfahl, A.; Rauh, D.; Binder, M.; Deibel, C.; Dyakonov, V. S-shaped current-voltage characteristics of organic solar devices. Phys. Rev. B 2010, 82, 115306. [Google Scholar] [CrossRef]
  36. Kim, J.; Kim, H.; Kim, G.; Back, H.; Lee, K. Soluble transition metal oxide/polymeric acid composites for efficient hole-transport layers in polymer solar cells. ACS Appl. Mater. Interfaces 2014, 6, 951–957. [Google Scholar] [CrossRef] [PubMed]
  37. Talledo, A.; Granqvist, C.G. Electrochromic vanadium-pentoxide-based films—Structural, electrochemical, and optical-properties. J. Appl. Phys. 1995, 77, 4655–4666. [Google Scholar] [CrossRef]
  38. Bullot, J.; Cordier, P.; Gallais, O.; Gauthier, M.; Babonneau, F. Thin-layers deposited from V2O5 gels 2. An optical-absorption study. J. Non Cryst. Solids 1984, 68, 135–146. [Google Scholar] [CrossRef]
  39. Meyer, J.; Zilberberg, K.; Riedl, T.; Kahn, A. Electronic structure of vanadium pentoxide: An efficient hole injector for organic electronic materials. J. Appl. Phys. 2011, 110, 033710–033715. [Google Scholar] [CrossRef]
  40. Hancox, I.; Rochford, L.A.; Clare, D.; Walker, M.; Mudd, J.J.; Sullivan, P.; Schumann, S.; McConville, C.F.; Jones, T.S. Optimization of a high work function solution processed vanadium oxide hole-extracting layer for small molecule and polymer organic photovoltaic cells. J. Phys. Chem. C 2013, 117, 49–57. [Google Scholar] [CrossRef]
  41. Negreira, A.S.; Aboud, S.; Wilcox, J. Surface reactivity of V2O5(001): Effects of vacancies, protonation, hydroxylation, and chlorination. Phys. Rev. B 2011, 83. [Google Scholar] [CrossRef]
  42. Silversmit, G.; Depla, D.; Poelman, H.; Marin, G.B.; de Gryse, R. Determination of the V2p XPS binding energies for different vanadium oxidation states (V5+ to V0+). J. Electron Spectrosc. Relat. Phenom. 2004, 135, 167–175. [Google Scholar] [CrossRef]
  43. Sawatzky, G.A.; Post, D. X-ray photoelectron and auger-spectroscopy study of some vanadium-oxides. Phys. Rev. B 1979, 20, 1546–1555. [Google Scholar] [CrossRef]
  44. Demeter, M.; Neumann, M.; Reichelt, W. Mixed-valence vanadium oxides studied by XPS. Surf. Sci. 2000, 454–456, 41–44. [Google Scholar] [CrossRef]
  45. Coulston, G.W.; Thompson, E.A.; Herron, N. Characterization of vpo catalysts by X-ray photoelectron spectroscopy. J. Catal. 1996, 163, 122–129. [Google Scholar] [CrossRef]
  46. Suchorski, Y.; Rihko-Struckmann, L.; Klose, F.; Ye, Y.; Alandjiyska, M.; Sundmacher, K.; Weiss, H. Evolution of oxidation states in vanadium-based catalysts under conventional XPS conditions. Appl. Surf. Sci. 2005, 249, 231–237. [Google Scholar] [CrossRef]
  47. Hermann, K.; Witko, M.; Druzinic, R.; Tokarz, R. Hydrogen assisted oxygen desorption from the V2O5(010) surface. Top. Catal. 2000, 11, 67–75. [Google Scholar] [CrossRef]
  48. Toledano, D.S.; Henrich, V.E.; Metcalf, P. Surface reduction of Cr-V2O3 by Co. J. Vacuum Sci. Technol. A 2000, 18, 1906–1914. [Google Scholar] [CrossRef]
  49. Ganduglia-Pirovano, M.V.; Sauer, J. Stability of reduced V2O5(001) surfaces. Phys. Rev. B 2004, 70. [Google Scholar] [CrossRef]
  50. Yuan, N.Y.; Li, J.H.; Lin, C.L. Valence reduction process from sol-gel V2O5 to VO2 thin films. Appl. Surf. Sci. 2002, 191, 176–180. [Google Scholar]
  51. Surnev, S.; Ramsey, M.G.; Netzer, F.P. Vanadium oxide surface studies. Prog. Surf. Sci. 2003, 73, 117–165. [Google Scholar] [CrossRef]
  52. Bermudez, V.M.; Williams, R.T.; Long, J.P.; Reed, R.K.; Klein, P.H. Photoemission-study of hydrogen adsorption on vanadium dioxide near the semiconductor-metal phase-transition. Phys. Rev. B 1992, 45, 9266–9271. [Google Scholar] [CrossRef]
  53. Greiner, M.T.; Helander, M.G.; Tang, W.M.; Wang, Z.B.; Qiu, J.; Lu, Z.H. Universal energy-level alignment of molecules on metal oxides. Nat. Mater. 2012, 11, 76–81. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) The chemical structure of PFDT2BT-8; and (b) energy levels of electrodes, HELs, and active layer materials used in this work.
Figure 1. (a) The chemical structure of PFDT2BT-8; and (b) energy levels of electrodes, HELs, and active layer materials used in this work.
Materials 09 00235 g001
Figure 2. The current density-voltage characteristics of PFDT2BT-8:PC70BM based solar cell with (□) PEDOT:PSS, ( Materials 09 00235 i001) s-V2Ox, and ( Materials 09 00235 i002) MoO3.
Figure 2. The current density-voltage characteristics of PFDT2BT-8:PC70BM based solar cell with (□) PEDOT:PSS, ( Materials 09 00235 i001) s-V2Ox, and ( Materials 09 00235 i002) MoO3.
Materials 09 00235 g002
Figure 3. The current density-voltage characteristics of organic solar cell with () unannealed s-V2Ox interlayer and annealed at ( Materials 09 00235 i001) 100 °C, ( Materials 09 00235 i002) 200 °C, ( Materials 09 00235 i003) 300 °C, and ( Materials 09 00235 i004) 400 °C for 30 min.
Figure 3. The current density-voltage characteristics of organic solar cell with () unannealed s-V2Ox interlayer and annealed at ( Materials 09 00235 i001) 100 °C, ( Materials 09 00235 i002) 200 °C, ( Materials 09 00235 i003) 300 °C, and ( Materials 09 00235 i004) 400 °C for 30 min.
Materials 09 00235 g003
Figure 4. Optical transmission spectra of the s-V2Ox films with different Annealing temperature; () unannealed, ( Materials 09 00235 i001) 100 °C, ( Materials 09 00235 i002) 200 °C, ( Materials 09 00235 i003) 300 °C, and ( Materials 09 00235 i004) 400 °C. The insert shows the absorption coefficient (αhν)2 as a function of the photon energy.
Figure 4. Optical transmission spectra of the s-V2Ox films with different Annealing temperature; () unannealed, ( Materials 09 00235 i001) 100 °C, ( Materials 09 00235 i002) 200 °C, ( Materials 09 00235 i003) 300 °C, and ( Materials 09 00235 i004) 400 °C. The insert shows the absorption coefficient (αhν)2 as a function of the photon energy.
Materials 09 00235 g004
Figure 5. AFM topography (2 µm × 2 µm) of a 10 nm thick s-V2Ox layer deposited on top of Si with native oxide; (a) as deposited; (b) annealed at 200 °C; (c) annealed at 400 °C. The average grain size was 21.3 nm, 16.8 nm, and 19.6 nm, respectively.
Figure 5. AFM topography (2 µm × 2 µm) of a 10 nm thick s-V2Ox layer deposited on top of Si with native oxide; (a) as deposited; (b) annealed at 200 °C; (c) annealed at 400 °C. The average grain size was 21.3 nm, 16.8 nm, and 19.6 nm, respectively.
Materials 09 00235 g005
Figure 6. XPS spectra of s-V2Ox thin film deposited in air. (a) The solid line represents the experimental XPS spectra; the dashed lines are decomposed XPS; the binding energy of V5+ peaks are higher than V4+ peaks (b) Photoelectron spectra of (i) unannealed film and those annealed at (ii) 200 °C; (iii) 300 °C, and (iv) 400 °C.
Figure 6. XPS spectra of s-V2Ox thin film deposited in air. (a) The solid line represents the experimental XPS spectra; the dashed lines are decomposed XPS; the binding energy of V5+ peaks are higher than V4+ peaks (b) Photoelectron spectra of (i) unannealed film and those annealed at (ii) 200 °C; (iii) 300 °C, and (iv) 400 °C.
Materials 09 00235 g006
Figure 7. UPS measurements for V2Ox films Span coated on ITO in ambient conditions with no annealing and annealed, () unannealed, ( Materials 09 00235 i001) 200 °C, ( Materials 09 00235 i002) 300 °C, and ( Materials 09 00235 i003) 400 °C. (a) shows the secondary electron cut-off region; and (b) shows the expanded region near the Fermi level; the insert shows the density of gap states formed about 1 eV below the Fermi level.
Figure 7. UPS measurements for V2Ox films Span coated on ITO in ambient conditions with no annealing and annealed, () unannealed, ( Materials 09 00235 i001) 200 °C, ( Materials 09 00235 i002) 300 °C, and ( Materials 09 00235 i003) 400 °C. (a) shows the secondary electron cut-off region; and (b) shows the expanded region near the Fermi level; the insert shows the density of gap states formed about 1 eV below the Fermi level.
Materials 09 00235 g007
Table 1. Summary of solar cell parameters with s-V2Ox buffer layer annealed at different temperatures for different periods of time.
Table 1. Summary of solar cell parameters with s-V2Ox buffer layer annealed at different temperatures for different periods of time.
Ann. Temp.Maximum PCE (%)Average PCE(av) (%)Voc (V)Jsc (mA·cm−2)FF (%)Rs (Ω·cm2)Rsh (Ω·cm2)
Non6.05.8 ± 0.140.8610.0 ± 0.2267.1 ± 1.211.6 ± 0.91264 ± 91
1006.05.9 ± 0.140.8610.1 ± 0.1368.3 ± 1.710.5 ± 1.01282 ± 215
2006.35.9 ± 0.260.8710.2 ± 0.1467.1 ± 2.610.3 ± 0.71346 ± 246
3005.65.3 ± 0.280.839.9 ± 0.4164.3 ± 2.712.6 ± 1.71095 ± 238
4005.25.0 ± 0.170.8310.1 ± 0.1259.6 ± 0.917.6 ± 0.4933 ± 84
Table 2. Summary of solar cell parameters with s-V2Ox buffer layer annealed at different temperatures for different periods of time.
Table 2. Summary of solar cell parameters with s-V2Ox buffer layer annealed at different temperatures for different periods of time.
Annealing TemperatureV4+ Oxidation StateV5+ Oxidation StateWork Function (eV)Valence Band (eV)Eg (eV)
Unannealing15%85%5.262.532.5
200 °C22.5%77.5%5.222.533.78
300 °C27%73%5.232.502.68
400 °C33%67%5.192.502.42

Share and Cite

MDPI and ACS Style

Alsulami, A.; Griffin, J.; Alqurashi, R.; Yi, H.; Iraqi, A.; Lidzey, D.; Buckley, A. Thermally Stable Solution Processed Vanadium Oxide as a Hole Extraction Layer in Organic Solar Cells. Materials 2016, 9, 235. https://doi.org/10.3390/ma9040235

AMA Style

Alsulami A, Griffin J, Alqurashi R, Yi H, Iraqi A, Lidzey D, Buckley A. Thermally Stable Solution Processed Vanadium Oxide as a Hole Extraction Layer in Organic Solar Cells. Materials. 2016; 9(4):235. https://doi.org/10.3390/ma9040235

Chicago/Turabian Style

Alsulami, Abdullah, Jonathan Griffin, Rania Alqurashi, Hunan Yi, Ahmed Iraqi, David Lidzey, and Alastair Buckley. 2016. "Thermally Stable Solution Processed Vanadium Oxide as a Hole Extraction Layer in Organic Solar Cells" Materials 9, no. 4: 235. https://doi.org/10.3390/ma9040235

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