Ag2S Quantum Dot-Sensitized Solar Cells by First Principles: The

Sep 7, 2017 - (36) With Ag2S QDs, MBN and MBA ligands also produce a red shift (0.38 and 0.54 eV, respectively). The absorption spectra are clearly mo...
0 downloads 8 Views 2MB Size
Subscriber access provided by GRIFFITH UNIVERSITY

Article 2

AgS Quantum Dot Sensitized Solar Cells by FirstPrinciples: The Effect of Capping Ligands and Linkers Javier Amaya Suárez, José J Plata, Antonio M. Márquez, and Javier Fernández Sanz J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.7b07731 • Publication Date (Web): 07 Sep 2017 Downloaded from http://pubs.acs.org on September 9, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry A is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Ag2S Quantum Dot Sensitized Solar Cells by First-Principles: The Effect of Capping Ligands and Linkers Javier Amaya Su´arez,† Jose J. Plata,‡,† Antonio M. M´arquez,† and Javier Fern´andez Sanz∗,† †Departmento de Qu´ımica F´ısica, Universidad de Sevilla, 41012 Sevilla, Spain ‡Department of Mechanical Engineering and Materials Science, Duke University, Durham, North Carolina 27708, USA E-mail: [email protected]

ACS Paragon Plus Environment

1

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 23

Abstract Quantum dots solar cells, QDSCs, are one of the candidates for being a reliable alternative to fossil fuels. However, the well studied CdSe and CdTe-based QDSCs present variety of issues for their use in consumer-goods applications. Silver sulfide, Ag2 S, is a promising material, but poor efficiency has been reported for QDSCs based on this compound. The potential influence of each component of QDSCs is critical and key for the development of more efficient devices based on Ag2 S. In this work, density functional theory (DFT) calculations have been carried out in order to study the nature of the optoelectronic properties for an anatase-TiO2 (101) surface sensitized with different silver sulfide nanoclusters. We have demonstrated how is possible a deep tune of its electronic properties modifying the capping ligands and linkers to the surface. Finally, an analysis of the electron injection mechanism for this system is presented.

Introduction Solar energy is one of the cleanest renewable energies, however, first (silicon wafers) and second (thin film technology) generation photovoltaics present an upper limit on the conversion efficiency of only 33%. Nanostructured particles or quantum dots (QDs) are one of the candidates for being the third-generation solar cells because they combine high performance and low cost. 1,2 The efficiency of QDs solar cells (QDSCs) has drastically growth during the last years 3–9 putting them on par with dye sensitized solar cells (DSSCs) and bulk heterojunction (BHJ) photovoltaic cells. 10,11 Moreover, QDSCs present significant advantages compared with DSSCs such as the photostability of inorganic materials, high molar extinction coefficients or tunable energy gaps by controlling the QD size. Although CdSe and CdTe have been extensively used as QDs in QDSCs, 12 they present a variety of issues such as low abundance, high cost and toxicity 13 which are crucial for their use in consumer-goods applications. These concerns have led to the scientific community to look for alternative materials such as Cu2 S, 14–19 Ag2 S, 20–22 SnS, 23,24 Sb2 S3 , 25 and CuInS2 . 26 Ag2 S is a nontoxic semiconductor material which seems to be a good candidate with a band gap around 1 eV. 27 Recently, silver sulfide based materials have been proposed as one of the most promising alternatives for hot-carrier solar cells (HCSCs). 28,29 Lin et al. have measured long carrier cooling times in Ag2 S NPs that can potentially lead 29 to a efficiency around 33% in HCSCs. Silver sulfide has been also used as a light absorber in organic ACS Paragon Plus Environment

2

Page 3 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

BHJ solar cells with a maximum power conversion efficiency of 3.21%. 30 However, poor efficiency has been reported for QDSCs based on Ag2 S and TiO2 . 20–22 TiO2 nanotubes with a ZnO recombination barrier layer were synthesized by Chen et al. obtaining very low short-circuit current density, JSC (lower than 1 mA · cm−2 ). 20 Tubtimtae et al. achieved a higher performance for a Ag2 S-sensitized TiO2 solar cell with a JSC of 7.3 mA · cm−2 ) but with a low efficiency of 0.76%. 21 A Ag2 S-sensitized ZnO solar cell has been also reported, improving the JSC to 13.7 mA · cm−2 , but also with a low efficiency (0.49%) due to the lower open-circuit voltage, VOC , (100 mV) than the one in the cell based on TiO2 (about 350 mV). 22 Some authors have explained the poor performance of these systems despite its broad absorption spectrum by the alignment between the bands of the Ag2 S sensitizer and the oxide electrode. 31 To improve that, Ji et al. have synthesized Ag2 Sx Se1−x -sensitized TiO2 systems where the electron–hole recombination is minimized. 32 Despite the fact that silver sulfides QDs are potential candidates for substituting conventional QDSCs based on CdSe and PbS QDs, a deeper insight of the electronic structure of this system is needed to explain the previous poor results and improve its efficiency. The optimization of the performance of these devices remains as the main challenge. The different components (QD, the oxide, linkers and capping ligands) play an important role in the mechanisms that govern its performance and are the key to improve the efficiency of these devices. 33–39 However, to the best of our knowledge, there are not systematic studies about the effect of all these components to the electronic properties of the Ag2 S-TiO2 QDSC. In this article, we combine a methodology that has been proven to describe correctly the electronic structure of bulk Ag2 S, 40 with a bottom-up strategy for the modelling of the QDSC. 19 This study represents a step forward for the understanding of the electron transfer mechanism in the Ag2 S-TiO2 QDSCs and a progress to optimize and rationally design more efficient devices.

Computational details and model All the calculations were performed using the Vienna ab-initio simulation package code 41–43 with the projector-augmented wave method (PAW). 44,45 The energies were computed with the exchangecorrelation functional proposed by Perdew, Burke and Ernzerhof, PBE, 46 based on the generalized ACS Paragon Plus Environment

3

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment

Page 4 of 23

Page 5 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

layers were relaxed while last layer was fixed to its bulk positions.

Results and discussion Bare Ag2 S quantum dots Quantum dots are semiconductor nanoparticles made of a few hundreds or thousands of atoms that are small enough to show electronic confinement effects. Due to the difficulty of carrying out a first-principles study of this size, it is mandatory to look for models with a smaller size but able to qualitatively describe their properties. Two different strategies have been used to model the Ag2n Sn QDs. The structures of the smallest clusters (n = 1 − 12) were optimized taking as starting point the topology of Cu2n Sn clusters previously proposed by Dehnen et al. 52,53 The biggest clusters (n = 20 − 50), on the contrary, were modeled taking as reference the geometry of bulk α − Ag2 S. 54 From the bulk structure a sphere-like cluster is defined, modifying the cutoff radius to select the appropriate number of Ag2 S units. The optimized structures are depicted in Figure S1.

Figure 2: Ag2 S QD energy (EAg2 S ) per Ag2 S unit. The horizontal line represents the energy per Ag2 S unit of bulk α − Ag2 S The relative stability of QDs is represented in Figure 2. It can be observed as the bigger the cluster, the higher the stability, which is connected to the number of not fully coordinated atoms. The values in Figure 2 indicate that energyPlus perEnvironment Ag2 S unit starts to converge around 6 units and ACSthe Paragon 5

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment

Page 6 of 23

Page 7 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Due to the high computational cost that such a long aliphatic chain would include in our model, we have selected a group of alternative thiols: Methanethiol (MT), 3-mercaptopropionic acid (MPA), ben-

zenethiol (BT), 4-(dimethylamino)benzenethiol (DAB), 4-methoxybenzenethiol (MBT), 4-mercaptobenzonitril (MBN) and 4-mercaptobenzoic acid (MBA) (see Figure 3). MT is the aliphatic thiol with the shortest chain and MPA has been widely synthesized and used as a linker molecule. 38,60 The rest of the ligands are aromatic molecules in which we have included different functional groups. These groups can present electron-donating and withdrawing effects and modify the optoelectronic properties of the QD stabilizing charges or holes that are produced when the electronic excitation occurs. 61,62 Each Ag atom of the clusters was coordinated to the thiol group of a ligand and the geometries were fully relaxed (see Figure 4). The HOMO-LUMO gap values obtained from the DOS calculations are shown in Table 1 as well as the energies for the first excitations in the absorption spectra (see Figure 5). The QD-ligands systems yields a band gap reduction of 0.08 - 0.38 eV due to the stabilization of the anti-bonding states. This stabilization also produces a red shift (0.52 - 1.05 eV) of the first absorption peaks, Eabs , in the absorption spectra compared to the bare quantum dots (see Figure S3). Different behaviors can be observed in absorption spectra depending on the nature of the chain. While Eabs keeps approximately constant for QDs saturated with aliphatic thiols regardless the carbon chain size (see Figure 5 (a)), these excitation energies highly depend on the substituents when aromatic ligands are involved (see Figure 5 (b)). Electron-donating groups, like −OCH3 or −N(CH3 )2 , tend to stabilize the photogenerated hole, which results in a red shift of the spectra. 62 In DAB and MBT, the electron-donating groups reduce the first excitation energy by 0.20 eV and 0.09 eV respectively. However, there is not a clear trend for ligands with electron-withdrawing groups (-CN, -COOH). For instance, 4-aminobenzonitrile (ABN) ligands shift the maxima to higher energies when coordinated to Cu2 S QDs, but a shift to lower energies was found when MBN was used. 19 The same tend was Table 1: Band gap energy , Eg , and first absorption peak energies, Eabs , for the saturated cluster with the different ligands. Energy values in eV. Ligand MT MPA BT DAB

Eg 2.37 2.35 2.28 2.19

Ligand MBT MBN MBA

Eabs 2.66 2.70 2.64 2.44

Eg 2.23 2.22 2.07

ACS Paragon Plus Environment

7

Eabs 2.55 2.26 2.17

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 23

found by Nadler et al. for the (CdSe)13 (ABN)6 system. 36 With Ag2 S QDs, MBN and MBA ligands also produce a red shift (0.38 and 0.54 eV respectively). The absorption spectra is clearly modified by the nature of the ligands but the origin of this modification is unclear. Bader charges of the QDs were calculated to analyze the impact of the charge transfer between the QD and the capping ligands on the optical properties (see Figure 6). There is a general trend in which the lower the negative charge of the cluster, the larger the red shift of the spectra. While in the aliphatic thiols, the lone pair of electrons of the sulfur atoms is strongly coordinated to the Ag atom of the QD, in the aromatic thiols, the benzene ring competes for these electrons. This competition can be tuned with electron-donating or withdrawing groups. MBA and MBN ligands present electron-withdrawing groups that reduce the electron donation of the thiol to the QD. However, ligands with electron donating groups such as DAB and MBT increase the charge transfer with respect the BT molecule. To demonstrate the complex and high impact of the ligands in the electronic structure of the QD, the absorption spectra of QD-MBA was calculated using three MBA-isomers (ortho, meta, and para). As it can be seen in Figure 7, there are strong modifications of the spectra depending on the isomer. The largest red shift is obtained for the o-MBA ligand and the smaller for the p-MBA molecules. These findings are in agreement with the experimental results obtained for MBA isomers adsorbed on Cu QDs, where the o-MBA ligand presents the largest shift to lower energies in the spectra. 63

Linkers and adsorption on TiO2 anatase (101) There are a different methods to link the nanoparticles on the oxide surface such as drop cast/spin cast, chemical bath and surface ionic layer deposition adsorption and reaction (SILAR). 1 Pre-synthesized QDs can also be adsorbed on the surface through an organic molecule that acts like an anchor which is known as the linker. This method shows some advantages over other methods because the number of attached QDs can be controlled as well as the linker can tailor the charge separation, recombination and transport mechanism. 56 For this study, three different linkers were selected as models. Cysteine (Cys) is a natural, abundant and inexpensive amino acid, which has been proven as excellent candidate not only theoretically but also experimentally. 16,19,56 The 3-mercaptopropionic acid (MPA) is frequently used for this purpose, for instance, attaching CdSe/CdTe core/shell nanoparticles on TiO2 . 60 As an ACS Paragon Plus Environment

8

Page 9 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Table 2: First, second and third absorption energy, Eabs , values for the isolated QD-MT and QD-MT-linker-TiO2 systems with different linkers. Energy values in eV. Linker Cys (c1) Cys (c2) MPA MBA QD

1st 2.59 2.66 2.65 2.54 2.65

2nd 2.65 2.76 2.73 2.59 2.74

3rd 2.71 2.84 2.83 2.65 2.83

aromatic linker, we have chosen 4-mercaptobenzoic acid (MBA), which has been reported to accelerate the electronic injection process with respect to non-aromatics linkers. 38 The molecular structures for these three molecules are shown in Figure 8. The adsorption of the linker molecules on the anatase surface is the next step in the construction of the model. Different sites and configurations were tested. For cysteine, the lower energy conformation corresponds to deprotonation and interaction through the carboxylic oxygen atoms with two surface Ti cations. 19 MPA or MBA molecules do not present an amino group, so the most stable configuration is the adsorption through the deprotonated carboxylic group (see Figure 9). In all cases, the linkers are adsorbed strongly to the surface with adsorption energies greater than 1 eV. To study the QD-surface interaction, all the silver atoms of the “61 ” cluster were coordinated with MT molecules except one that was coordinated to the linker anchored to the anatase surface. MT was selected as capping ligand because of the computational cost associated to the size of the system (370 atoms). The adsorption of one QD over the anatase surface creates a QD monolayer with a distance between images of around 5 ˚ A (see Figure 10 (a)). The electronic structure of the semiconductor is strongly modified by the QD adsorption. While bare TiO2 shows a band gap value around 3 eV, the gap is almost closed when the QD is adsorbed. As it can be observed in Figure 10 (b), the QD states fill the anatase band gap. However, it seems that changing the linker does not affect significantly the electronic structure of the system. This severe modification of the electronic structure affects the absorption spectrum of the TiO2 . Due to its band gap value, anatase shows a wide band beyond 3 eV in the optical spectrum, however, new absorption peaks appear at lower energies when QD is adsorbed (Figure 10 (c)). Table 2 shows the position of the first three peaks in the absorption spectra using different linkers. The different values obtained for each linker demonstrate ACS Paragon Plus Environment

9

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 23

the importance of this molecule not only as a anchor between the QD and the surface but also as a possible target to tune the electronic properties of the system. It is worth to mention that also the adsorption configuration of the linker can modify the electronic structure. While Cys(c1) induces a red shift in the spectra, Cys(c2) shows almost the same energy for the first absorption peak compared to the isolated QD/ligands system. MBA is the linker that produce the most significant shift to lower energies being in agreement with the effect that was observed when it was used as a capping ligand. It is possible to get information about the electron-injection mechanism comparing the position of the optical absorption maxima of the QD before and after being adsorbed on the surface. In a direct mechanism, electrons are excited from the QD HOMO state to the edge of the semiconductor conduction band. This transition requires a lower energy than the HOMO-LUMO transition energy so the absorption peak appears at a lower energy than the first peak of the spectra of the isolated QD/ligand system. On the other hand, the indirect mechanism means that electrons are excited from the QD HOMO state to the QD LUMO state. Then, the electron is injected into the semiconductor conduction band, since the QD LUMO state is overlapping with the conduction band. For this reason, the position of the optical maxima remains almost unaltered for an indirect mechanism. A schematic representation of these mechanisms is shown in Figure 10 (d). As it is shown in Figure 10 (c) and Table 2, the position of most the QD peaks remain almost unaltered especially for Cys (c2) and MPA. This fact points to a mainly indirect electron-injection mechanism. However, there is a shift of the first peaks to lower energies for Cys (c1) and MBA, which indicate some contribution of a direct mechanism too, especially for the MBA linker.

Conclusions The effect of ligands, linkers and support oxide on the behavior of Ag2 S-based QDSCs was examined using first principles DFT calculations. The model was build using a bottom-up approach to differentiate the effect of each element on the optoelectronic properties of the system. While there are not significant changes in the spectra using aliphatic ligands, there is a common shift to lower energies for most of the aromatic ligands. The effect of the ligand can be highly tuned even selecting different isomers. It was found that o-MBA presents a higher red shift in the spectra than m-MBA and p-MBA. ACS Paragon Plus Environment

10

Page 11 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

It seems that the linker selection does not severely modify the system’s electronic structure even selecting linkers that strongly interact with the anatase (101) TiO2 surface. The QD/ligands states are localized filling the band gap of the anatase, sensitizing the visible zone of the spectrum where the bare TiO2 does not show any absorption feature. Considering the position of the QDs peaks in the spectra before and after its adsorption on the anatase surface, a mainly indirect injection mechanism is proposed. However, a small contribution of a direct mechanism is observed notably when the MBA is used as linker.

Acknowledgement This work was funded by the Ministerio de Econom´ıa y Competitividad (Spain), the EU FEDER program, and the Junta de Andaluc´ıa, Grants CTQ2015-64669-P and P12-FQM-1595.

Supporting Information Available • QDs geometries and DOS and absorption spectra of “61 ” QD.

References (1) Kamat, P. V. Quantum Dot Solar Cells. The Next Big Thing in Photovoltaics. J. Phys. Chem. Lett. 2013, 4, 908–918. (2) R¨ uhle, S.; Shalom, M.; Zaban, A. Quantum-Dot-Sensitized Solar Cells. ChemPhysChem 2010, 11, 2290–2304. (3) Ip, A.; Thon, S.; Hoogland, S.; Voznyy, O.; Zhitomirsky, D.; Debnath, R.; Levina, L.; Rollny, L.; Carey, G.; Fischer, A. et al. Hybrid Passivated Colloidal Quantum Dot Solids. Nat. Nanotechnol. 2012, 7, 577–582. (4) Kim, H.-S.; Lee, C.-R.; Im, J.-H.; Lee, K.-B.; Moehl, T.; Marchioro, A.; Moon, S.-J.; Humphry-

ACS Paragon Plus Environment

11

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 23

Baker, R.; Yum, J.-H.; Moser, J. et al. Lead Iodide Perovskite Sensitized All-Solid-State Submicron Thin Film Mesoscopic Solar Cell with Efficiency Exceeding 9%. Sci. Rep. 2012, 2, 591. (5) Lee, M.; Teuscher, J.; Miyasaka, T.; Murakami, T.; Snaith, H. Efficient Hybrid Solar Cells Based on Meso-Superstructured Organometal Halide Perovskites. Science 2012, 338, 643–647. (6) Jiao, S.; Du, J.; Du, Z.; Long, D.; Jiang, W.; Pan, Z.; Li, Y.; Zhong, X. Nitrogen-Doped Mesoporous Carbons as Counter Electrodes in Quantum Dot Sensitized Solar Cells with a Conversion Efficiency Exceeding 12%. J. Phys. Chem. Lett. 2017, 8, 559–564. (7) Du, J.; Du, Z.; Hu, J.-S.; Pan, Z.; Shen, Q.; Sun, J.; Long, D.; Dong, H.; Sun, L.; Zhong, X. et al. Zn-Cu-In-Se Quantum Dot Solar Cells with a Certified Power Conversion Efficiency of 11.6%. J. Am. Chem. Soc. 2016, 138, 4201–4209. (8) Zhao, K.; Pan, Z.; Mora-Ser´o, I.; C´anovas, E.; Wang, H.; Song, Y.; Gong, X.; Wang, J.; Bonn, M.; Bisquert, J. et al. Boosting Power Conversion Efficiencies of Quantum-Dot-Sensitized Solar Cells Beyond 8% by Recombination Control. J. Am. Chem. Soc. 2015, 137, 5602–5609. (9) Pan, Z.; Mora-Ser´o, I.; Shen, Q.; Zhang, H.; Li, Y.; Zhao, K.; Wang, J.; Zhong, X.; Bisquert, J. High-Efficiency ’Green’ Quantum Dot Solar Cells. J. Am. Chem. Soc. 2014, 136, 9203–9210. (10) Yella, A.; Lee, H.-W.; Tsao, H. N.; Yi, C.; Chandiran, A. K.; Nazeeruddin, M.; Diau, E. W.G.; Yeh, C.-Y.; Zakeeruddin, S. M.; Gr¨atzel, M. Porphyrin-Sensitized Solar Cells with Cobalt (II/III)-Based Redox Electrolyte Exceed 12% Efficiency. Science 2011, 334, 629–634. (11) He, F.; Yu, L. How Far Can Polymer Solar Cells Go? In Need of a Synergistic Approach. J. Phys. Chem. Lett. 2011, 2, 3102–3113. (12) Robel, I.; Subramanian, V.; Kuno, M.; Kamat, P. V. Quantum Dot Solar Cells. Harvesting Light Energy with CdSe Nanocrystals Molecularly Linked to Mesoscopic TiO2 Films. J. Am. Chem. Soc. 2006, 128, 2385–2393. (13) Zhao, Y.; Burda, C. Development of Plasmonic Semiconductor Nanomaterials with Copper Chalcogenides for a Future with Sustainable Energy Materials. Energy Environ. Sci. 2012, 5, 5564–5576.

ACS Paragon Plus Environment

12

Page 13 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(14) Kim, C. S.; Choi, S. H.; Bang, J. H. New Insight into Copper Sulfide Electrocatalysts for Quantum Dot-Sensitized Solar Cells: Composition-Dependent Electrocatalytic Activity and Stability. ACS Appl. Mater. Interfaces 2014, 6, 22078–22087. (15) Ye, M.; Wen, X.; Zhang, N.; Guo, W.; Liu, X.; Lin, C. In Situ Growth of CuS and Cu1.8 S Nanosheet Arrays as Efficient Counter Electrodes for Quantum Dot-Sensitized Solar Cells. J. Mater. Chem. A 2015, 3, 9595–9600. (16) Ratanatawanate, C.; Bui, A.; Vu, K.; Balkus, K. J. Low-Temperature Synthesis of Copper(II) Sulfide Quantum Dot Decorated TiO2 Nanotubes and Their Photocatalytic Properties. J. Phys. Chem. C 2011, 115, 6175–6180. (17) Nelwamondo, S.; Moloto, M.; Krause, R.; Moloto, N. Synthesis and Characterization of AlanineCapped Water Soluble Copper Sulphide Quantum Dots. Mater. Lett. 2012, 75, 161–164. (18) Mousavi-Kamazani, M.; Zarghami, Z.; Salavati-Niasari, M. Facile and Novel Chemical Synthesis, Characterization, and Formation Mechanism of Copper Sulfide (Cu2 S, Cu2 S/CuS, CuS) Nanostructures for Increasing the Efficiency of Solar Cells. J. Phys. Chem. C 2016, 120, 2096–2108. (19) Amaya Su´arez, J.; Plata, J. J.; M´arquez, A. M.; Fdez. Sanz, J. Effects of Capping Ligands, Linkers and Oxide Surface in the Electron Injection Mechanism of Copper Sulfide Quantum Dots Sensitized Solar Cells. Phys. Chem. Chem. Phys. 2017, (20) Chen, C.; Xie, Y.; Ali, G.; Yoo, S. H.; Cho, S. O. Improved Conversion Efficiency of Ag2 S Quantum Dot-Sensitized Solar Cells Based on TiO2 Nanotubes with a ZnO Recombination Barrier Layer. Nanoscale Res. Lett. 2011, 6, 462. (21) Tubtimtae, A.; Wu, K.-L.; Tung, H.-Y.; Lee, M.-W.; Wang, G. J. Ag2 S Quantum Dot-Sensitized Solar Cells. Electrochem. Commun. 2010, 12, 1158 – 1160. (22) Wu, J.-J.; Chang, R.-C.; Chen, D.-W.; Wu, C.-T. Visible to Near-Infrared Light Harvesting in Ag2 S Nanoparticles/ZnO Nanowire Array Photoanodes. Nanoscale 2012, 4, 1368–1372. (23) Xu, Y.; Al-Salim, N.; Bumby, C. W.; Tilley, R. D. Synthesis of SnS Quantum Dots. J. Am. Chem. Soc. 2009, 131, 15990–15991. ACS Paragon Plus Environment 13

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 23

(24) Liu, H.; Liu, Y.; Wang, Z.; He, P. Facile Synthesis of Monodisperse, Size-Tunable SnS Nanoparticles Potentially for Solar Cell Energy Conversion. Nanotechnology 2010, 21, 105707. (25) Itzhaik, Y.; Niitsoo, O.; Page, M.; Hodes, G. Sb2 S3 -Sensitized Nanoporous TiO2 Solar Cells. J. Phys. Chem. C 2009, 113, 4254–4256. (26) Kuo, K.-T.; Liu, D.-M.; Chen, S.-Y.; Lin, C.-C. Core-Shell CuInS2 /ZnS Quantum Dots Assembled on Short ZnO Nanowires with Enhanced Photo-Conversion Efficiency. J. Mater. Chem. 2009, 19, 6780–6788. (27) Rodr´ıguez, A. N.; Nair, M. T. S.; Nair, P. K. Structural, Optical and Electrical Properties of Chemically Deposited Silver Sulfide Thin Films. Semicond. Sci. Technol. 2005, 20, 576. (28) Lin, S.; Feng, Y.; Wen, X.; Zhang, P.; Woo, S.; Shrestha, S.; Conibeer, G.; Huang, S. Theoretical and Experimental Investigation of the Electronic Structure and Quantum Confinement of WetChemistry Synthesized Ag2 S Nanocrystals. J. Phys. Chem. C 2015, 119, 867–872. (29) Lin, S.; Feng, Y.; Wen, X.; Harada, T.; Kee, T. W.; Huang, S.; Shrestha, S.; Conibeer, G. Observation of Hot Carriers Existing in Ag2 S Nanoparticles and Its Implication on Solar Cell Application. J. Phys. Chem. C 2016, 120, 10199–10205. (30) Chen, C.; Zhai, Y.; Li, F.; Yue, G. Fabrication of Silver Sulfide Thin Films for Efficient Organic Solar Cells with High Short-Circuit Currents Based on Double Heterojunctions. J. Power Sources 2015, 298, 259 – 268. (31) Shen, H.; Jiao, X.; Oron, D.; Li, J.; Lin, H. Efficient Electron Injection in Non-Toxic Silver Sulfide Ag2 S Sensitized Solar Cells. J. Power Sources 2013, 240, 8 – 13. (32) Ji, C.; Zhang, Y.; Zhang, X.; Wang, P.; Shen, H.; Gao, W.; Wang, Y.; Yu, W. W. Synthesis and Characterization of Ag2 Sx Se1−x Nanocrystals and their Photoelectrochemical Property. Nanotechnology 2017, 28, 065602. (33) Grandhi, G. K.; M., A.; Viswanatha, R. Understanding the Role of Surface Capping Ligands in Passivating the Quantum Dots Using Copper Dopants as Internal Sensor. J. Phys. Chem. C 2016, 120, 19785–19795.

ACS Paragon Plus Environment

14

Page 15 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(34) Kilina, S.; Ivanov, S.; Tretiak, S. Effect of Surface Ligands on Optical and Electronic Spectra of Semiconductor Nanoclusters. J. Am. Chem. Soc. 2009, 131, 7717–7726. (35) Inerbaev, T. M.; Masunov, A. E.; Khondaker, S. I.; Dobrinescu, A.; Plamadˇa, A.-V.; Kawazoe, Y. Quantum Chemistry of Quantum Dots: Effects of Ligands and Oxidation. J. Chem. Phys. 2009, 131, 044106. (36) Nadler, R.; Sanz, J. F. Effect of Capping Ligands and TiO2 Supporting on the Optical Properties of a (CdSe)13 Cluster. J. Phys. Chem. A 2015, 119, 1218–1227. (37) Mora-Ser´o, I.; Gim´enez, S.; Moehl, T.; Fabregat-Santiago, F.; Lana-Villareal, T.; G´omez, R.; Bisquert, J. Factors Determining the Photovoltaic Performance of a CdSe Quantum Dot Sensitized Solar Cell: the Role of the Linker Molecule and of the Counter Electrode. Nanotechnology 2008, 19, 424007. (38) Wang, H.; McNellis, E. R.; Kinge, S.; Bonn, M.; C´anovas, E. Tuning Electron Transfer Rates through Molecular Bridges in Quantum Dot Sensitized Oxides. Nano Lett. 2013, 13, 5311–5315. (39) Yang, J.; Oshima, T.; Yindeesuk, W.; Pan, Z.; Zhong, X.; Shen, Q. Influence of Linker Molecules on Interfacial Electron Transfer and Photovoltaic Performance of Quantum Dot Sensitized Solar Cells. J. Mater. Chem. A 2014, 2, 20882–20888. (40) Amaya Su´arez, J.; Plata, J. J.; M´arquez, A. M.; Sanz, J. F. Structural, Electronic and Optical Properties of Copper, Silver and Gold Sulfide: a DFT Study. Theor. Chem. Acc. 2016, 135, 70. (41) Kresse, G.; Furthm¨ uller, J. Efficient Iterative Schemes for Ab initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B 1996, 54, 11169–11186. (42) Kresse, G.; Furthm¨ uller, J. Efficiency of Ab-initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comput. Mater. Sci. 1996, 6, 15–50. (43) Kresse, G.; Hafner, J. Ab Initio Molecular Dynamics for Liquid Metals. Phys. Rev. B 1993, 47, 558–561. ACS Paragon Plus Environment

15

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 23

(44) Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector Augmented-Wave Method. Phys. Rev. B 1999, 59, 1758–1775. (45) Bl¨ochl, P. Projector Augmented-Wave Method. Phys. Rev. B 1994, 50, 17953–17979. (46) Perdew, J.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. (47) Dudarev, S.; Botton, G.; Savrasov, S.; Humphreys, C.; Sutton, A. Electron-Energy-Loss Spectra and the Structural Stability of Nickel Oxide: An LSDA+U Study. Phys. Rev. B 1998, 57, 1505–1509. (48) Park, J. B.; Graciani, J.; Evans, J.; Stacchiola, D.; Ma, S.; Liu, P.; Nambu, A.; Sanz, J. F.; Hrbek, J.; Rodriguez, J. A. High Catalytic Activity of Au/CeOx /TiO2 (110) Controlled by the Nature of the Mixed-Metal Oxide at the Nanometer Level. Proc. Natl. Acad. Sci. 2009, 106, 4975–4980. (49) Plata, J. J.; Collico, V.; M´arquez, A. M.; Sanz, J. F. Understanding Acetaldehyde Thermal Chemistry on the TiO2 (110) Rutile Surface: From Adsorption to Reactivity. J. Phys. Chem. C 2011, 115, 2819–2825. (50) Harris, J. Simplified Method for Calculating the Energy of Weakly Interacting Fragments. Phys. Rev. B 1985, 31, 1770–1779. (51) Gajdoˇs, M.; Hummer, K.; Kresse, G.; Furthm¨ uller, J.; Bechstedt, F. Linear Optical Properties in the Projector-Augmented Wave Methodology. Phys. Rev. B 2006, 73, 045112. (52) Dehnen, S.; Sch¨afer, A.; Ahlrichs, R.; Fenske, D. An Ab-initio Study of Structures and Energetics of Copper Sulfide Clusters. Chem. Eur. J 1996, 2, 429–435. (53) Dehnen, S.; Fenske, D. [Cu24 S12 (PMeiPr2 )12 ], [Cu28 S14 (PtBu2 Me)12 ], [Cu50 S25 (PtBu2 Me)16 ], [Cu70 Se35 (PtBu2 Me)21 ], [Cu31 Se15 (SeSiMe3 )(PtBu2 Me)12 ] and [Cu48 Se24 (PMe2 Ph)20 ]: Sulfur- and Selenium-Bridged Copper Clusters. Chem. Eur. J. 1996, 2, 1407–1416. ACS Paragon Plus Environment

16

New

Page 17 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(54) Sadanaga, R.; Sueno, S. X-Ray Study of the Alpha-beta Transition of Ag2S. Miner. J. Jpn. 1967, 5, 124–143. (55) Vogel, R.; Hoyer, P.; Weller, H. Quantum-Sized PbS, CdS, Ag2 S, Sb2 S3 , and Bi2 S3 Particles as Sensitizers for Various Nanoporous Wide-Bandgap Semiconductors. J. Phys. Chem. 1994, 98, 3183–3188. (56) Margraf, J. T.; Ruland, A.; Sgobba, V.; Guldi, D. M.; Clark, T. Quantum-Dot-Sensitized Solar Cells: Understanding Linker Molecules through Theory and Experiment. Langmuir 2013, 29, 2434–2438. (57) Zhang, Y.; Hong, G.; Zhang, Y.; Chen, G.; Li, F.; Dai, H.; Wang, Q. Ag2 S Quantum Dot: A Bright and Biocompatible Fluorescent Nanoprobe in the Second Near-Infrared Window. ACS Nano 2012, 6, 3695–3702. (58) Zhuang, Z.; Peng, Q.; Wang, X.; Li, Y. Tetrahedral Colloidal Crystals of Ag2 S Nanocrystals. Angew. Chem. Int. Ed. 2007, 46, 8174–8177. (59) Zhang, Y.; Liu, Y.; Li, C.; Chen, X.; Wang, Q. Controlled Synthesis of Ag2 S Quantum Dots and Experimental Determination of the Exciton Bohr Radius. J. Phys. Chem. C 2014, 118, 4918–4923. (60) Wang, J.; Mora-Ser´o, I.; Pan, Z.; Zhao, K.; Zhang, H.; Feng, Y.; Yang, G.; Zhong, X.; Bisquert, J. Core/Shell Colloidal Quantum Dot Exciplex States for the Development of Highly Efficient Quantum-Dot-Sensitized Solar Cells. J. Am. Chem. Soc. 2013, 135, 15913–15922. (61) Tan, Y.; Jin, S.; Hamers, R. J. Influence of Hole-Sequestering Ligands on the Photostability of CdSe Quantum Dots. J. Phys. Chem. C 2013, 117, 313–320. (62) Tan, Y.; Jin, S.; Hamers, R. J. Photostability of CdSe Quantum Dots Functionalized with Aromatic Dithiocarbamate Ligands. ACS Appl. Mater. Interfaces 2013, 5, 12975–12983. (63) Lin, Y.-J.; Chen, P.-C.; Yuan, Z.; Ma, J.-Y.; Chang, H.-T. The Isomeric Effect of Mercaptobenzoic Acids on the Preparation and Fluorescence Properties of Copper Nanoclusters. Chem. Commun. 2015, 51, 11983–11986. ACS Paragon Plus Environment 17

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment

Page 18 of 23

Page 19 of 23

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment

Page 20 of 23

Page 21 of 23

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 8: Structure of the linkers: mercaptobenzoic acid (MBA).

Page 22 of 23

cysteine (Cys), 3-mercaptopropionic acid (MPA) and 4-

Figure 9: MPA adsorption geometry onto anatase (101). Colors: Ti, cyan; O, red; C, black; H, white; S, yellow.

ACS Paragon Plus Environment

22

Page 23 of 23

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment