Next Article in Journal
Construction of ZnCdS Quantum-Dot-Modified CeO2 (0D–2D) Heterojunction for Enhancing Photocatalytic CO2 Reduction and Mechanism Insight
Next Article in Special Issue
Insight into the Structural and Performance Correlation of Photocatalytic TiO2/Cu Composite Films Prepared by Magnetron Sputtering Method
Previous Article in Journal
Engineering Xylose Isomerase for Industrial Applications
Previous Article in Special Issue
V6O13 Micro-Flower Arrays Grown In Situ on Ni Foam as Efficient Electrocatalysts for Hydrogen Evolution at Large Current Densities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Variations in Power Conversion Efficiency on n-Type Dye-Sensitized Solar Cells with Synthesized TiO2 Nanoparticle: A Thickness Effect of Active Layer

1
Convergence Research Center for Energy and Environmental Science, Sungkyunkwan University, Suwon 16419, Republic of Korea
2
Institute of Basic Science, Sungkyunkwan University, Suwon 16419, Republic of Korea
3
Department of Chemistry, Sungkyunkwan University, Suwon 16419, Republic of Korea
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(9), 598; https://doi.org/10.3390/catal14090598
Submission received: 31 July 2024 / Revised: 2 September 2024 / Accepted: 3 September 2024 / Published: 6 September 2024

Abstract

:
Recently, many researchers have made progress in studies aimed at enhancing the power conversion efficiency (PCE) of n-type dye-sensitized solar cells (DSSCs). This paper presents a systematic investigation focused on improving the PCEs of n-type DSSCs by synthesizing TiO2 nanoparticle active layers and varying their thickness. The study found that increasing the TiO2 layer thickness up to 17 µm resulted in a steady 41% increase in PCE, primarily owing to the enhanced photocurrent density in the n-type DSSCs. This improvement is attributed to the enhanced light harvesting effect. As a result, n-type DSSC with 17 µm thick TiO2 layer demonstrates a relatively high Jsc value of 8.42 mA/cm2, achieving an overall PCE of 4.02%. In contrast, the n-type DSSC with a 6 µm thick TiO2 layer exhibits a much lower Jsc value of 5.55 mA/cm2, leading to a reduced PCE. This result represents at least a 52% increase in current density, indicating that the optimal thickness of the TiO2 active layer is a critical factor influencing the PCE of DSSCs.

1. Introduction

Since the dye-sensitized solar cell (DSSC) was first developed in 1991 with an initial efficiency of 7.12% [1], considerable efforts have been made to improve its power conversion efficiency (PCE) [2,3,4,5,6,7,8]. Despite considerable advancements over the past two decades, the highest efficiencies reported remain at approximately 12% [2,3,7], and achieving efficiencies above 10% is uncommon without highly optimized fabrication conditions. Given the considerable advantages of DSSCs, such as their manufacturing cost being only one-fifth that of conventional silicon solar cells [1,2,3], and their versatility in applications, including the production of richly colored and transparent products, the potential of DSSCs remains substantial. The research results will remain valuable. To enhance the efficiency of DSSCs, various techniques have been proposed, including the deposition of a thin tunneling barrier layer on the substrate [9,10] or the oxide surface [11], co-sensitization using different dyes [12], and post-treatment with a TiCl4 precursor [13,14]. Additionally, surface texturing has been widely adopted because it confines the incident light in the electrode, thereby increasing the photocurrent density [15].
To date, titanium oxide (TiO2), a well-known photocatalytic semiconductor, has typically been used as the active layer in the working electrode of DSSCs owing to its relatively large band gap and suitable conduction band energy. It is well established that achieving high PCE requires the energy band structure of TiO2 to be well-aligned with that of the dye. Specifically, if the conduction band energy of TiO2 is higher than the LUMO energy of the dye, electron injection from the dye becomes challenging. Given the properties of commonly used ruthenium-based dyes (N3, N719), the selection of compatible oxides is extremely limited. This makes TiO2 an ideal material because its conduction band energy is approximately 0.2 eV lower than the LUMO of dyes, facilitating efficient electron injection. Additionally, TiO2 is well-suited for use as a scattering layer owing to its chemical stability and dye absorption capability. As a result, many DSSCs are constructed usingTiO2 nanoparticle films with a TiO2 scattering layer on top of the active layer. For instance, Grätzel et al. especially used a mesoporous TiO2 film layer to achieve a 1000-fold increase in surface area [1,16]. By incorporating a scattering layer, the overall PCE can be further enhanced.
The thickness of the TiO2 active layer is also a crucial factor influencing the PCE of DSSCs. As the thickness increases, the surface area of the TiO2 layer expands, leading to greater dye adsorption [17,18,19]. However, this also lengthens the electron diffusion time, which can compete with the electron lifetime, thereby impairing electron transport properties. Additionally, the increased thickness reduces light transmittance, preventing the concentration of photoelectrons from increasing in proportion to the increased dye adsorption [20]. If we can control both surface area and thickness of the TiO2 active layer by changing crystal structure, particle size, and morphology, a high PCE will be achieved. For this purpose, our research group has synthesized TiO2 nanoparticles, nanowires and nanorods continuously for the last 15 years [21,22,23], even though we did not obtain the result expected. Despite this situation, in this paper, we aim to present a systematic study of the impact of TiO2 active layer thickness on the PCE of n-type DSSCs. To achieve this, we synthesized other types of TiO2 nanoparticles using the hydrothermal method and examined the effect of their thickness on the PCE of the fabricated n-type DSSCs. The performance was then compared to that of DSSC based on commercially available P-25 TiO2.

2. Results and Discussion

The solar energy power (P) can be calculated using the following equation:
P = IV
The energy conversion efficiency (η) of DSSCs is defined as the ratio of the maximum electrical power output (Pmax) to the solar energy power input (P). Pmax is calculated by multiplying the maximum current (Jmp) by the maximum voltage (Vmp). The solar energy power input is determined as the product of the irradiance of the incident light, measured in W/m2, according to the following equation:
η = Pmax/P = Jmp × Vmp/P
The ideal power density of DSSCs is a product of the maximum short-circuit current Jsc and maximum open-circuit voltage Voc. The fill factor (ff) is a measure of the quality of the solar cell, and it can be calculated by comparing Jsc and Voc to Jmp and Vmp by the following equation:
ff = (Jmp × Vmp)/(Jsc × Voc)
Therefore, we can obtain the power conversion efficiency (η) of DSSCs by the following equation.
η = Pmax/P = (Jsc × Voc × ff)/P
where P (100 mW/cm2) is the solar cell testing standard under terrestrial conditions with air mass AM1.5.
According to Equation (4), achieving high ƞ in DSSCs requires maximizing both Jsc and Voc, as well as ff. Thus, researchers have focused on integrating metal oxide nanostructures into DSSCs to increase dye adsorption and thereby enhance the photocurrent (Jsc).The fundamental operating principle of DSSCs mirrors that of photocatalytic reactions, where the generation of hole–electron pairs is crucial for both high PCE and efficient catalysis. Among various metal oxide materials, TiO2 nanostructures are particularly favored for DSSC and photocatalysis applications owing to their stability and low production cost.
Figure 1 shows a schematic diagram of the doctor-blade method used for coating TiO2 active layers. The thickness of the active layer can be precisely controlled by adjusting both the number of 3M tape layers and the number of doctor-blade coating cycles. Using these techniques, the TiO2 active layers were accurately fabricated in a thickness range of 6–24 µm. Detailed procedures for the fabrication and characterization of TiO2 based DSSCs have been previously published [21,22].
Figure 2a presents a typical XRD pattern of TiO2 nanoparticles synthesized and annealed at 700 °C. The prominent diffraction peak at 25° corresponds to the (101) rutile plane, indicating that the primary crystal growth direction is along the {101} direction. A smaller diffraction peak at 38° is attributed to the (101) plane of anatase phase. Additionally, several other peaks corresponding to either (200) and (211) rutile or (211) anatase crystallites are also observed. By analyzing the peak areas (and considering their FWHM values), we calculated the crystalline phase ratio of rutile to anatase, which is very similar to that of P-25, with values of 78% rutile and 22% anatase. Park et al. reported that anatase crystals generally have smaller particle sizes compared to rutile crystals, which results in a larger surface area for anatase. Consequently, more dye molecules can be adsorbed on the anatase surface, leading to improved light absorption and increased photocurrent. Additionally, the interactions between particles in anatase crystals are stronger than those in rutile crystals, which enhance electron mobility and boost photocurrent [24]. Conversely, rutile crystals are more stable and more effective at scattering light compared to anatase crystals, and finding the optimal combination of these phases remains a crucial research objective. Figure 2b shows an atomic force microscopy (AFM) image of the same film depicted in Figure 2a, with a scan size of 10 × 10 µm2. The image reveals craters approximately 0.5 µm in size on the surface of the fluorine-doped tin oxide (FTO) glass substrate, along with small nanosized clusters [see the enlarged image in Figure 2b]. The average surface roughness was measured at 0.274 nm, indicating a relatively larger surface area compared to a film coated with commercially available P-25 powder. This rough surface increases surface haze owing to a decrease in particle size. Therefore, this rough surface can produce more electrons in photovoltaic cells, resulting in the increase of photocurrent (Jsc) from our synthesized TiO2 nanoparticles.
Figure 3 presents (a) scanning electron microscopy (SEM) and (b) electron dispersion spectroscopy (EDS) data for the TiO2 nanoparticles synthesized after calcining at 450 °C and annealing at 700 °C. As shown in Figure 3a, the SEM image reveals spherical nanoparticles with some aggregate morphologies and an average particle size of 25 nm (see SEM inset). To determine the composition ratio of the synthesized TiO2 nanoparticles, we measured the EDS spectrum. Figure 3b shows the EDS spectrum along with elemental mapping data (inset), which shows peaks corresponding only to Ti and O, with a weight percent ratio of Ti to O at 49.88:50.12. This indicates that a slightly excess of oxygen likely attributed to water adsorption or further oxidation in the air exhibits. Considering the differences in atomic sensitivity in EDS, the atomic percent indicates 75% O and 25% Ti, which falls within acceptable error limits. Using theseTiO2 nanoparticles, we fabricated the active layers of n-type DSSCs with varying film thicknesses by adjusting the number of 3M tape layers and doctor-blade coating cycles, ranging from 1 to 5 layers.
Figure 4 shows the SEM surface morphologies (a–c) and cross-sectional images (d–g) of TiO2 nanoparticle film layers with varying numbers of doctor-blade coatings such as 1 layer (d), 2 layer (e), 3 layer (f), and 5 layer (g), respectively. By adjusting both the numbers of 3M tape layers and doctor-blade coating cycle, we controlled the thickness of the TiO2 nanoparticle film layers. The thickness of the TiO2 active layers varied as follows: 6 µm (1 layer), 12 µm (2 layers), 17 µm (3 layers), and 24 µm (5 layers) with an increasing number of doctor-blade coating cycles. As the thickness increased, the surface roughness also increased, indicating an increase in surface area (Figure 4a–c). When we fabricated films thicker than 17 µm (3 layers), both optical transmittance and PCE of the n-type DSSCs decreased, largely owing to minimal changes in dye absorption [see Figure 5b as well]. Therefore, in this study, we excluded TiO2 nanoparticle films thicker than 17 µm (3 layers). The cross-sectional images (d–g) clearly illustrate the variations in TiO2 nanoparticle film thickness with different numbers of doctor-blade coating cycles. Based on this study, it is evident that, while light efficiency increases with thickness up to a certain point, it subsequently decreases when the thickness exceeds this optimal level. This decrease is likely attributable to increased electron diffusion time and reduced light transmittance, as described earlier. As the layer thickness increases, however, the distance that electrons generated by the dye must travel to reach the electrode becomes too great, which can lead to recombination and reduced electron injection efficiency into the electrode.
Figure 5a shows UV-Visible absorption spectra measured after N719 dye adsorption on the TiO2 active layers, as shown in Figure 4. The baseline UV-Visible absorption spectrum was recorded from a bare FTO glass substrate without dye adsorption. The intensity of the dye adsorption peaks (as well as peak area), particularly around 500 nm, which correspond to the π–π* transition in the N719 dye molecule, increases with the thickness of the TiO2 active layers. This suggests that a thicker TiO2 layer, with its larger surface area, allows for greater dye adsorption. Using Beer’s law, we calculated the relative amounts of adsorbed dye on the TiO2 active layers based on the data presented in Figure 5a. Figure 5b shows the variation in thickness (left, black color) and dye adsorption amounts (right, red color) with the number of doctor-blade coating cycles. As the number of coating cycles increases, both thickness and dye adsorption amounts increase from 1.2 × 10−7 mol/cm3 (6 µm, 1 layer) to 2.9 × 10−7 mol/cm3 (24 µm, 5 layers). However, the rate of increase begins to level off after three coating cycles. This indicates that the amount of dye molecules increases with the thickness of the active layers up to a critical thickness. Additionally, we observed that the PCE of DSSCs with active layer thicknesses exceeding 17 µm (3 layer) did not improve and even declined at 24 µm (see Table 1). This suggests that, as the TiO2 active layer becomes thicker, the distance electrons generated by the dye must travel to reach the electrode becomes too long. Consequently, recombination may occur, preventing efficient electron injection into the electrode. Therefore, achieving high surface roughness (i.e., a large surface area) combined with an optimal TiO2 active layer thickness can absorb more dye molecules, enhancing the light-trapping or scattering effect, particularly in the long-wavelength region. This leads to an increase in the PCE of the photovoltaic device. This result indicates that TiO2 nanoparticle-based DSSCs are highly effective in enhancing light-trapping or light-scattering properties in photovoltaic cells by adjusting thickness and surface area. However, it is important to note that, while nanoparticle films demonstrate considerably higher PCE compared to other nanostructures, they also face a clear disadvantage in charge transport owing to the excessive number of adsorbed dyes.
The photocurrent (Jsc) and photovoltage (Voc) of the solar cell devices, with an active area of 0.25 cm2, were measured using simulated sunlight at AM-1.5 produced by solar simulator. Figure 6 shows the J–V curves for TiO2 nanoparticle film-based DSSCs with varying active layer thicknesses. The photovoltaic characteristics of these n-type DSSC devices are summarized in Table 1. As illustrated in Figure 6 and detailed in Table 1, the TiO2 nanoparticle film-based DSSCs with active layer thicknesses of less than 12 µm (2 layers) exhibit considerably lower PCE values compared to the commercially available P-25 based DSSC, which has a PCE of 3.77%. This decrease is likely attributable to the relatively low crystalline quality (poor polycrystallinity) of the synthesized TiO2 nanoparticles. Despite the lower PCE values compared to P-25 based DSSCs, the experiment shows that Jsc gradually increases with the thickness of TiO2 nanoparticle active layers, from 1 to 3 layers (see the y-axis changes in Figure 6). This increases in Jsc results in an increase in PCE from 2.86% to 4.02%, which surpasses the P-25 based DSSC. In the case of 5 layer, however, the PCE was dropped to 3.91% due mostly to the decrease of Jsc and ff rather than Voc. This indicates that if a thickness of active layer has over optimum value, the distance electrons generated by the dye should long travel to reach the electrode owing to relatively prolonged recombination lifetime (This will be confirmed later). From the Table 1, we concluded that the considerable increase in PCE by 41% is primarily attributed to a large improvement in photocurrent density (at least 52%) in the n-type DSSCs, although the values of Voc and ff remained relatively constant. Ohno et al. reported the analysis results of composition ratio between rutile and anatase crystalline phase, as well as average particle sizes with commercially available P-25 (Degussa, Frankfurt, Germany) TiO2 powder of 70–80% anatase and 20–30% rutile, with average particle sizes ranging from 20 to 30 nm, respectively [25]. Since our synthesized TiO2 nanoparticles have a similar composition ratio (78% anatase and 22% rutile, with little high amount of anatase due possibly to relatively low annealing temperature [26]) and average particle size (25 nm) with P-25, they can be comparable to each other even though we obtained a relatively low PCE value.
To understand this result in detail, we performed an analysis using electrochemical impedance spectroscopy (EIS). Figure 7a shows the Nyquist plots of the EIS data for the same DSSCs shown in Figure 6. The EIS data reveal four components: Z1 (high-frequency region, 100 kHz–500 Hz), Z2 (middle-frequency region, 500 Hz–1 Hz), Z3 (low-frequency region, 1 Hz–100 mHz), and the Ohmic series resistance of the TCO (Rh). In Figure 7a, three distinct semicircles are observed in the measured frequency range of 100 mHz–100 kHz. Rh in the high-frequency region is associated with the resistance of the electrolyte and the FTO, while the impedances (or resistances) Z1, Z2, and Z3 correspond to different aspects of the charge transfer and diffusion processes. Z1 is the charge transfer interface resistance between the Pt counter electrode and the electrolyte (R1), Z2 is the charge transfer resistance at the TiO2/electrolyte interface (R2), and Z3 is the Warburg diffusion resistance in the electrolyte (R3) [27]. As we can see in the Figure 7b, however, in the analysis of a traditional equivalent circuit model for DSSCs, a constant photo-generated current source, a series parasitic resistance (i.e., series resistance, Rs) and a parallel parasitic resistance (i.e., parallel resistance, Rp) are generally included. For an understanding of the electronic behaviors of the solar cell, therefore, two types of resistance values, series and parallel resistance, were identified to measure the change in charge-carrier mobility with the energy level of the transport layer due to changes inside the solar cell. Because internal impedance is inversely proportional to solar cell performance, DSSCs with thicker TiO2 active layers exhibit relatively lower impedance compared to those with thinner TiO2 layers [see the maximum values of each second semicircle in Figure 7a]. Thus, our data in Figure 7a confirm a decreasing trend in all impedances with increasing thickness of the TiO2 active layer. Among the resistances R1, R2, and R3, the R2 value (32.13 Ω) for the DSSC with a thick TiO2 active layer of 17 µm (3 layers) showed lower impedance compared to those of DSSCs with thinnerTiO2 layers (see also Table 2). However, the EIS data of thickness shows an abnormal trend for the relatively high second semicircle and low impedance compared with other data. This suggests a longer electron diffusion time, indicating that the DSSC with the thickerTiO2 layer below critical value can has a higher electron density and reduced electron recombination, which improves charge transport. Consequently, the DSSC with the 17 µm thick TiO2 active layer demonstrates the highest photocurrent (8.42 mA/cm2) and the best cell performance (η = 4.02%) in this study. However, the PCE begins to decline when the thickness exceeds 17 µm (3 layers) owing to extended electron diffusion times and reduced light transmittance in thicker layers of lower-quality TiO2 nanoparticles. This indicates the need for further experiments to enhance light harvesting through improved light-trapping or light-scattering techniques. Exploring alternative active layers, such as 1-dimensional TiO2 nanostructures, may offer better performance compared to nanoparticles.
In order to understand in detail the main reason for the thickness effect on the Jsc and PCE, calculations of electron lifetime (τe) and recombination lifetime (τr) were carried out as following procedures. The electron diffusion length (Ln) is determined by the diffusion coefficient (De) and the electron lifetime (τe) [16].
Ln = Deτe
According to Equation (5), we can realize that the Ln becomes longer as the τe increases. The electron lifetime (τe) is calculated by Equation (6) using the peak frequency (fp,max) of the second semicircle from the Nyquist diagram shown in Figure 7a [28].
τe = 1/fp,max
The peak frequency (fp,max) of the second semicircle and the electron lifetime (τe) for the different TiO2 thicknesses are shown in Table 2, together with the values of R2 and recombination lifetime (τr). The peak frequency (fp,max) decreases and the electron lifetime (τe) increases as the thickness of TiO2 increases. Based on our obtained EIS data shown in Figure 7a, relative impedance values are obtained from the DSSCs with different thickness of TiO2 active layers [29]. We could also obtain a value of electron diffusion time (τe) theoretically by using Equation (6) and its results are summarized in Table 2. The obtained τe values are 5.24 × 10−3 (6 µm), 5.71 × 10−3 (12 µm), 6.04 × 10−3 (17 µm), and 5.53 × 10−3 s (24 µm), respectively. The increase in the τe makes the electron diffuse and transfer more easily due to the increase in the Ln, since the τr is closely related to both electron lifetime (τe) and electron diffusion length (Ln). Recombination lifetimes (τr) were therefore determined from the photovoltage (Voc) decay curve (see Figure 8) and realized a tendency to oppose between τe and τr (see Table 2). As shown in Figure 7a (see also Table 1) and Figure 8, the highest efficiency cell showed the lowest τr owing to lowest recombination. Therefore, in this work, the Jsc and PCE are increased as the TiO2 thickness increases up to 17 μm because the internal resistance (R2) related to the electron transport in the TiO2/dye/electrolyte interface is decreased and the electron lifetime (τe) is increased, reflecting a very short electron recombination time (τr). The performance of the DSSC with a thickness above 17 μm, however, is decreased because the recombination of the electron is strengthened. Finally, it is demonstrated that the optimal TiO2 thickness is 17 μm for the best performance of the DSSC.

3. Materials and Methods

To prepareTiO2 nanoparticles suitable for use as an active layer in DSSCs, we used a hydrothermal method, synthesizing the TiO2 nanoparticles using a titanium isopropoxide solution in a temperature range 50–200 °C. The nanoparticles were then calcined and annealed at 450 °C and 700 °C, respectively. Before calcining and annealing, we measured the average particle sizes, which were initially in the range of 15–20 nm. After calcining and annealing, the particle sizes increased to 25–30 nm. The annealing process aimed to achieve a crystalline phase ratio similar to that of P-25 (Degussa) standard TiO2 material, which has a crystalline phase ratio of, e.g., 70–80% anatase and 20–30% rutile.
To prepare the TiO2 paste, we combined 0.5 g of standard P-25 (Degussa) TiO2 powder with the annealed synthesized TiO2 nanoparticles. This mixture was ground with 1 mL of water containing acetylacetone (Aldrich, Saint Louis, MO, USA) and 0.25 g of hydroxypropyl cellulose (Aldrich) to prevent particle re-aggregation. The resulting paste was then spread uniformly and evenly applied to the surface of FTO-coated glass substrates using the doctor-blade technique to fabricate DSSC devices. After application, all doctor-bladed film samples were annealed in air at 450 °C for 60 min to remove the solvent. To fabricate the DSSC devices, the annealed, doctor-bladed films were sensitized by immersing them in an ethanol solution containing 0.5 mM N719 dye (Solaronix Inc., Aubonne, Switzerland) for 24 h. After sensitization, the films were sandwiched between a platinum-coated FTO counter electrode and the sensitized film. The electrodes were separated by a 20 µm thick polypropylene spacer, and the internal cell space was filled with a liquid electrolyte (Solaronix Inc., Iodolyte AN-50). The active area of each DSSC was 0.25 cm2. To observe the typical characteristics of fabricated n-type DSSC devices, photocurrent density–voltage (I–V) curves (SUN 2000) were measured using a xenon lamp under AM 1.5 filter at 100 mW/cm2 illumination by a source meter (Model 2400, Keithley Instrument, Inc., Solon, OH, USA) under open circuit conditions. The analysis of electrochemical impedance spectroscopy (EIS) was then performed with the electrochemical analyzer (Biologic science instrument, Seyssinet-Pariset, France, SP-150) in the range of the frequency from 100 mHz to 100 kHz under AC voltage with a perturbation amplitude of 10 mV applied in the EIS measurement. In order to the light absorption of dye molecules with different TiO2 thicknesses, UV/VIS spectra are measured from the wavelength of 400 nm to the wavelength of 800 nm by means of the UV/VIS spectrophotometer (OPTIZEN 3220 UV, Mecasys, Daejeon, Republic of Korea).

4. Conclusions

We successfully synthesized TiO2 nanoparticles with a composition ratio and crystallinity similar to those of commercially available P-25 TiO2 powder. Subsequently, we fabricated TiO2 nanoparticle-based n-type DSSCs with varying active layer thicknesses and compared their photovoltaic properties to those of P-25 TiO2-based DSSCs. The results clearly demonstrate the impact of TiO2 layer thickness on n-type DSSC performance. As the thickness of the TiO2 layer increased up to 17 µm, the PCE gradually improved from 2.86% to 4.02%, primarily owing to the increased photocurrent density in the n-type DSSCs. Notably, the DSSC with a 17 µm thick TiO2 active layer achieved a maximum Jsc of 8.42 mA/cm2, which is considerably higher compared to the 5.55 mA/cm2 obtained with the 6 µm thick TiO2 active layer, resulting in a current density increase of at least 52%. However, the relatively poor crystallinity and smaller particle size of the synthesized TiO2 active layers adversely affected both Voc and ff. Therefore, further research is essential to address these issues.

Author Contributions

Conceptualization, S.-H.N. and J.-H.B.; methodology, S.-H.N. and J.-H.B.; software, D.-W.J.; validation, S.-H.N., D.-W.J. and J.-H.B.; formal analysis, S.-H.N. and D.-W.J.; investigation, S.-H.N. and D.-W.J.; resources, S.-H.N. and D.-W.J.; data curation, S.-H.N. and D.-W.J.; writing—original draft preparation, S.-H.N.; writing—review and editing, J.-H.B.; visualization, D.-W.J.; supervision, J.-H.B.; project administration, J.-H.B.; funding acquisition, J.-H.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (No. NRF-2019R1A6A1A10073079).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. O’Regan, B.; Grätzel, M. A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films. Nature 1991, 353, 737–740. [Google Scholar] [CrossRef]
  2. Grätzel, M. Photoelectrochemical cells. Nature 2001, 414, 338–344. [Google Scholar] [CrossRef] [PubMed]
  3. Nazeeruddin, M.K.; Péchy, P.; Renouard, T.; Zakeeruddin, S.M.; Humphry-Baker, R.; Comte, P.; Liska, P.; Cevey, L.; Costa, E.; Shklover, V.; et al. Engineering of efficient panchromatic sensitizers for nanocrystalline TiO2-based solar cells. J. Am. Chem. Soc. 2001, 123, 1613–1624. [Google Scholar] [CrossRef]
  4. Grätzel, M. Conversion of sunlight to electric power by nanocrystalline dye-sensitized solar cells. J. Photochem. Photobiol. A Chem. 2004, 164, 3–14. [Google Scholar] [CrossRef]
  5. Lee, K.; Park, S.W.; Ko, M.J.; Kim, K.; Park, N.-G. Selective positioning of organic dyes in a mesoporous inorganic oxide film. Nat. Mater. 2009, 8, 665–671. [Google Scholar] [CrossRef]
  6. Chiba, Y.; Islam, A.; Watanabe, Y.; Komiya, R.; Koide, N.; Han, L. Dye-Sensitized Solar Cells with Conversion Efficiency of 11.1%. Jpn. J. Appl. Phys. 2006, 45, L638. [Google Scholar] [CrossRef]
  7. Wei, M.; Konishi, Y.; Zhou, H.; Yanagida, M.; Sugihara, H.; Arakawa, H. Highly efficient dye-sensitized solar cells composed of mesoporous titanium dioxide. J. Mater. Chem. 2006, 16, 1287–1293. [Google Scholar] [CrossRef]
  8. Cameron, P.J.; Peter, L.M. Characterization of titanium dioxide blocking layers in dye-sensitized nanocrystalline solar cells. J. Phys. Chem. B 2003, 107, 14394–14400. [Google Scholar] [CrossRef]
  9. Xia, J.; Masaki, N.; Jiang, K.; Wada, Y.; Yanagida, S. Importance of blocking layers at conducting glass/TiO2 interfaces in dye-sensitized ionic-liquid solar cells. Chem. Lett. 2006, 35, 252–253. [Google Scholar] [CrossRef]
  10. Palomares, E.; Clifford, J.N.; Haque, S.A.; Lutz, T.; Durrant, J.R. Control of charge recombination dynamics in dye sensitized solar cells by the use of conformally deposited metal oxide blocking layers. J. Am. Chem. Soc. 2003, 125, 475–482. [Google Scholar] [CrossRef]
  11. Clifford, J.N.; Palomares, E.; Nazeeruddin, M.K.; Thampi, R.; Grätzel, M.; Durrant, J.R. Multistep electron transfer processes on dye co-sensitized nanocrystalline TiO2 films. J. Am. Chem. Soc. 2004, 126, 5670–5671. [Google Scholar] [CrossRef] [PubMed]
  12. Park, N.-G.; Schlichthörl, G.; van de Lagemaat, J.; Cheong, H.M.; Mascarenhas, A.; Frank, A.J. Dye-sensitized TiO2 solar cells: Structural and photoelectrochemical characterization of nanocrystalline electrodes formed from the hydrolysis of TiCl4. J. Phys. Chem. B 1999, 103, 3308–3314. [Google Scholar] [CrossRef]
  13. Sommeling, P.M.; O’Regan, B.C.; Haswell, R.R.; Smit, H.J.P.; Bakker, N.J.; Smits, J.J.T.; Kroon, J.M.; van Roosmalen, J.A.M. Influence of a TiCl4 post-treatment on nanocrystalline TiO2 films in dye-sensitized solar cells. J. Phys. Chem. B 2006, 110, 19191–19197. [Google Scholar] [CrossRef]
  14. Wang, Z.-S.; Kawauchi, H.; Kashima, T.; Arakawa, H. Significant influence of TiO2 photoelectrode morphology on the energy conversion efficiency of N719 dye-sensitized solar cell. Coord. Chem. Rev. 2004, 248, 1381–1389. [Google Scholar] [CrossRef]
  15. Ito, S.; Zakeeruddin, S.M.; Humphry-Baker, R.; Liska, P.; Charvet, R.; Comte, P.; Nazeeruddin, M.K.; Péchy, P.; Takata, M.; Miura, H.; et al. High-Efficiency Organic-Dye-Sensitized Solar Cells Controlled by Nanocrystalline-TiO2 Electrode Thickness. Adv. Mater. 2006, 18, 1202–1205. [Google Scholar] [CrossRef]
  16. Grätzel, M. Solar Energy Conversion by Dye-Sensitized Photovoltaic Cells. Inorg. Chem. 2005, 44, 6841–6851. [Google Scholar] [CrossRef]
  17. Abdel-Galeil, M.M.; Kumar, R.; Matsuda, A.; El-Shater, R.E. Investigation on influence of thickness variation effect of TiO2 film, spacer and counter electrode for improved dye-sensitized solar cells performance. Optik 2021, 227, 166108. [Google Scholar] [CrossRef]
  18. Zhang, H.; Wang, W.; Liu, H.; Wang, R.; Chen, Y.; Wang, Z. Effects of TiO2 film thickness on photovoltaic properties of dye-sensitized solar cell and its enhanced performance by graphene combination. Mater. Res. Bull. 2014, 49, 126–131. [Google Scholar] [CrossRef]
  19. Kao, M.C.; Chen, H.Z.; Young, S.L.; Kung, C.Y.; Lin, C.C. The effects of the thickness of TiO2 films on the performance of dye-sensitized solar cells. Thin Solid Film. 2009, 517, 5096–5099. [Google Scholar] [CrossRef]
  20. Nakade, S.; Matsuda, M.; Kambe, S.; Saito, Y.; Kitamura, T.; Sakata, T.; Wada, Y.; Mori, H.; Yanagida, S. Dependence of TiO2 Nanoparticle Preparation Methods and Annealing Temperature on the Efficiency of Dye-Sensitized Solar Cells. J. Phys. Chem. B 2002, 106, 10004–10010. [Google Scholar] [CrossRef]
  21. Nam, S.-H.; Ju, D.-W.; Boo, J.-H. Comparison of Photovoltaic Properties of TiO2 Electrodes Prepared with Nanoparticles and Nanorods. J. Nanosci. Nanotechnol. 2014, 14, 9406–9410. [Google Scholar] [CrossRef] [PubMed]
  22. Nam, S.-H.; Boo, J.-H. Effect of TiO2 layer thickness on the energy conversion efficiency of dye-sensitized solar cells. Funct. Mater. Lett. 2022, 15, 2251034. [Google Scholar] [CrossRef]
  23. Nam, S.H.; Hyun, J.-S.; Boo, J.-H. Synthesis of TiO2 thin films using single molecular precursors by MOCVD method for dye-sensitized solar cells application and study on film growth mechanism. Mater. Res. Bull. 2012, 47, 2717–2721. [Google Scholar] [CrossRef]
  24. Park, N.-G.; van de Lagemaat, J.; Frank, A.J. Comparison of dye-sensitized rutile- and anatase-based TiO2 solar cells. J. Phys. Chem. B 2000, 104, 8989–8994. [Google Scholar] [CrossRef]
  25. Ohno, T.; Sarukawa, K.; Tokieda, K.; Matsumura, M. Morphology of a TiO2 Photocatalyst (Degussa, P-25) Consisting of Anatase and Rutile Crystalline Phases. J. Catal. 2001, 203, 82–86. [Google Scholar] [CrossRef]
  26. Ohtani, B.; Prieto-Mahaney, O.O.; Li, D.; Abe, R. What is Degussa (Evonik) P25? Crystalline composition analysis, reconstruction from isolated pure particles and photocatalytic activity test. J. Photochem. Photobiol. A Chem. 2010, 216, 179–182. [Google Scholar] [CrossRef]
  27. Kim, T.-Y.; Song, D.; Barea, E.M.; Lee, J.H.; Kim, Y.R.; Cho, W.; Lee, S.; Rahman, M.M.; Bisquert, J.; Kang, Y.S. Origin of high open-circuit voltage in solid state dye-sensitized solar cells employing polymer electrolyte. Nano Energy 2016, 28, 455–461. [Google Scholar] [CrossRef]
  28. Hoshikawa, T.; Ikebe, T.; Kikuchi, R.; Eguchi, K. Effects of electrolyte in dye-sensitized solar cells and evaluation by impedance spectroscopy. Electrochim. Acta 2006, 51, 5286–5294. [Google Scholar] [CrossRef]
  29. Shin, I.; Seo, H.; Son, M.-K.; Kim, J.-K.; Prabakar, K.; Kim, H.-J. Analysis of TiO2 thickness effect on characteristic of a dye-sensitized solar cell by using electrochemical impedance spectroscopy. Curr. Appl. Phys. 2010, 10, S422–S424. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the doctor-blade method for coating TiO2 active layers.
Figure 1. Schematic diagram of the doctor-blade method for coating TiO2 active layers.
Catalysts 14 00598 g001
Figure 2. (a) XRD pattern and (b) AFM image of a synthesized TiO2 nanoparticle obtained after calcining at 450 °C and annealing at 700 °C.
Figure 2. (a) XRD pattern and (b) AFM image of a synthesized TiO2 nanoparticle obtained after calcining at 450 °C and annealing at 700 °C.
Catalysts 14 00598 g002
Figure 3. (a) SEM and (b) EDS data for a synthesized TiO2 nanoparticle obtained after calcining at 450 °C and annealing at 700 °C.
Figure 3. (a) SEM and (b) EDS data for a synthesized TiO2 nanoparticle obtained after calcining at 450 °C and annealing at 700 °C.
Catalysts 14 00598 g003
Figure 4. SEM morphologies (ac) and cross-sectional views (dg) of TiO2 nanoparticle film layers with different numbers of doctor-blade coatings.
Figure 4. SEM morphologies (ac) and cross-sectional views (dg) of TiO2 nanoparticle film layers with different numbers of doctor-blade coatings.
Catalysts 14 00598 g004
Figure 5. (a) UV-Visible absorption spectra measured after N719 dye adsorption on the TiO2 active layers shown in Figure 4, and (b) the variation in thickness (black color) and dye adsorption amounts (red color) with the number of doctor-blade coating cycles. The baseline represents the UV-Visible absorption spectrum of the FTO glass substrate without dye adsorption.
Figure 5. (a) UV-Visible absorption spectra measured after N719 dye adsorption on the TiO2 active layers shown in Figure 4, and (b) the variation in thickness (black color) and dye adsorption amounts (red color) with the number of doctor-blade coating cycles. The baseline represents the UV-Visible absorption spectrum of the FTO glass substrate without dye adsorption.
Catalysts 14 00598 g005
Figure 6. Photocurrent density (J)–Voltage (V) curves for TiO2 nanoparticle film based DSSCs with different thicknesses.
Figure 6. Photocurrent density (J)–Voltage (V) curves for TiO2 nanoparticle film based DSSCs with different thicknesses.
Catalysts 14 00598 g006
Figure 7. (a) Nyquist plots of the electrochemical impedance spectroscopy (EIS) data obtained for the same DSSCs as in Figure 6. (b) A traditional equivalent circuit model utilized for the analysis of EIS data.
Figure 7. (a) Nyquist plots of the electrochemical impedance spectroscopy (EIS) data obtained for the same DSSCs as in Figure 6. (b) A traditional equivalent circuit model utilized for the analysis of EIS data.
Catalysts 14 00598 g007
Figure 8. Variations of Voc and Jsc as a function of TiO2 active layer thickness.
Figure 8. Variations of Voc and Jsc as a function of TiO2 active layer thickness.
Catalysts 14 00598 g008
Table 1. Photovoltaic characteristics of TiO2 nanoparticle-based DSSCs with different thicknesses.
Table 1. Photovoltaic characteristics of TiO2 nanoparticle-based DSSCs with different thicknesses.
Sample NamePCE (%)Voc (V)Jsc (mA/cm2)ff (%)
Base Cell (P-25) (0 µm)3.770.747.2969.6
TEST-6 1 layer (6 µm)2.860.735.5570.7
TEST-6 2 layer (12 µm)3.760.717.6768.9
TEST-6 3 layer (17 µm)4.020.708.4268.2
TEST-6 5 layer (24 µm)3.910.738.25 66.7
Table 2. DSSC parameters of internal resistance (R2), peak frequency (fp,max), electron lifetime (τe), and recombination lifetime (τr) extracted from EIS measurements [Figure 7a] and voltage decay curve (Figure 8).
Table 2. DSSC parameters of internal resistance (R2), peak frequency (fp,max), electron lifetime (τe), and recombination lifetime (τr) extracted from EIS measurements [Figure 7a] and voltage decay curve (Figure 8).
Sample NameR2 (Ω)fp,max (Hz)τe (ms)τr (s)
TEST-6 1 layer (6 µm)39.3519.085.241.38
TEST-6 2 layer (12 µm)35.6017.515.711.29
TEST-6 3 layer (17 µm)32.1316.576.040.92
TEST-6 5 layer (24 µm)30.5418.095.531.32
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

Nam, S.-H.; Ju, D.-W.; Boo, J.-H. Variations in Power Conversion Efficiency on n-Type Dye-Sensitized Solar Cells with Synthesized TiO2 Nanoparticle: A Thickness Effect of Active Layer. Catalysts 2024, 14, 598. https://doi.org/10.3390/catal14090598

AMA Style

Nam S-H, Ju D-W, Boo J-H. Variations in Power Conversion Efficiency on n-Type Dye-Sensitized Solar Cells with Synthesized TiO2 Nanoparticle: A Thickness Effect of Active Layer. Catalysts. 2024; 14(9):598. https://doi.org/10.3390/catal14090598

Chicago/Turabian Style

Nam, Sang-Hun, Dong-Woo Ju, and Jin-Hyo Boo. 2024. "Variations in Power Conversion Efficiency on n-Type Dye-Sensitized Solar Cells with Synthesized TiO2 Nanoparticle: A Thickness Effect of Active Layer" Catalysts 14, no. 9: 598. https://doi.org/10.3390/catal14090598

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