Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Optical, structural, and electrical properties of PEDOT:PSS thin films doped with silver nanoprisms

Open Access Open Access

Abstract

Conductive thin films of poly(3,4-ethylenedioxythiophene) poly(styrenesulfonate) (PEDOT:PSS) are common buffer layers widely applied in organic solar cells. In order to explore the implications of the localized surface plasmon resonance of noble metal nanoparticles in such applications, we prepared a series of hybrid PEDOT:PSS thin films doped with different proportions of colloidal Ag nanoprisms. Various characterization techniques, including transmission electron microscopy, regular and conductive atomic force microscopy, optical absorption, goniophotometry, and four-point probe resistance measurements, were applied to study the effects of Ag nanoprisms on the optical, structural, and electrical properties of the hybrid films. Through analyzing the Bidirectional Reflectance Distribution Functions (BRDF) of different hybrid films, we compared among different hybrid films the proportions of light being scattered and absorbed over various reflected angles. In terms of optical properties, with higher Ag nanoprism concentration, increased light scattering was found in the hybrid films which can potentially improve the light harvest in organic solar cells. In terms of structural and electrical properties, the surface roughness and the global sheet resistance of hybrid films were found to increase as the concentration of Ag nanoprism increases. The magnitude of the sheet resistance in these hybrid films was reduced to a level comparable to or smaller than those observed in pristine PEDOT:PSS films by increasing the extent of post-synthesis nanoprism purification and by applying organic solvent additives.

© 2014 Optical Society of America

1. Introduction

Over the last two decades, metal-polymer nanocomposites, metallic nanoparticles embedded in conducting polymer matrix, have attracted much attention due to their tunable magnetic, mechanical, electrical or optical properties [1–3]. These promising properties enable them to be applied in many applications such as sensors [4], light emitting diodes [5], solar cells [6] and capacitors [7]. Among various metal-polymer nanocomposites, there is significant interest in recent years on the properties of poly(3,4-ethylenedioxythiophene):poly(styrenesulfonate) (PEDOT:PSS) and their mixture with silver nanoparticles. PEDOT:PSS is a conductive polymer. Due to its suitable work function and beneficial effects in surface roughness [8,9], it is commonly used as a buffer layer between the anodic electrode and the organic photoactive layer in organic solar cells [10–12]. Instead of inserting sub-wavelength-dimension Ag nanoparticles (Ag NPs) inside the photoactive layer [13,14], their incorporation into PEDOT:PSS films has been found to enhance the absorption of the composite in the visible spectrum due to localized surface plasmon resonance (LSPR) [15–17]. In addition to absorption, light scattering [12,16] in this case was also found to increase which can be beneficial for light-trapping in the photoactive layer. It has also been reported that the incorporation of Ag NPs in PEDOT:PSS can effectively improve the conductivity of PEDOT:PSS [18–20] and enhance the surface roughness [6,12]. Combining all these beneficial effects in optical, structural and electrical properties, hybrid Ag-PEDOT:PSS thin films are of high potentials for organic solar cells towards better photovoltaic performance [6,12,21–23].

In this work we synthesized colloidal Ag nanoprisms (NPSMs) and fabricated a series of hybrid Ag NPSM-PEDOT:PSS thin films containing various amount of Ag NPSMs. Compared to spherical Ag NPs, Ag NPSMs exhibit LSPR with a wider wavelength tunability-window from the visible to the near-IR spectrum [24]. Despite this attractive LSPR tunability, the incorporation of Ag NPSMs in PEDOT:PSS and the resultant properties have not been explored so far. In this work a series of structural, optical, and electrical characterizations were performed on NPSM-PEDOT:PSS hybrid thin films by techniques including transmission electron microscopy (TEM), optical absorption, goniophotometry, regular and conductive atomic force microscopy (C-AFM), and four-point-probe resistance measurements. In particular, goniophotometry allows quantitative surface reflectance studies and the calculation of Bidirectional Reflectance Distribution Functions (BRDF) of sample surfaces, which are likely to be non-Lambertian diffusive surfaces [25]. Comparing the results from these characterization techniques, we study how the optical, structural and electrical properties of PEDOT:PSS thin films are modified by the addition of different amounts of Ag NPSMs, shedding light on the perspectives of their future applications in organic solar cells.

2. Experiment details

2.1 Ag NPSM synthesis

The Ag NPSMs used in this work were prepared by the seeded growth method from Aherne et al [26]. The seeds were synthesized by adding an aqueous silver nitrate (AgNO3, 5 mL, 0.5 mM) into a freshly prepared mixture of trisodium citrate (Na3C6H5O7·2H2O, 5mL, 2.5mM), poly(sodium styrenesulphonate) (PSS, 1,000 kDa, 0.25mL) and sodium borohydride (NaBH4, 0.3 mL, 10 mM) under rigorous stirring. The Ag NPSMs were produced by slowly adding AgNO3 (5 mL, 0.5 mM; 1 mL/min) into a mixed solution containing distilled water (5 mL), ascorbic acid (75 µL, 10 mM) and different quantities of the seed solution. Finally, another aqueous trisodium citrate solution (0.5 mL, 25 mM) was added into the above solution to stabilize Ag NPSMs. De-ionized (DI) water was used throughout the entire preparation.

2.2 Preparation of hybrid PEDOT:PSS-Ag NPSM solutions and films

The Ag NPSM solution (10 mL) was centrifuged at 6,000 rpm for 15 min by using tetrahydrofuran (THF, 35 mL) as the non-solvent. A pellet of Ag NPSM precipitate was obtained at the bottom of the centrifuge tube and the supernatant was removed. The pellets were re-dissolved in 0.5 mL DI water and then mixed with 0.5 mL PEDOT:PSS solution (Heraeus CleviosTM P) by ultrasonic bath. We then obtained a series of hybrid PEDOT:PSS-Ag NPSM solutions of different Ag NPSM concentration: 0.6 mg/mL (with 4 pellets of Ag NPSM precipitate), 1.6 mg/mL (10 pellets) and 2.6 mg/mL (16 pellets). The pure Ag NPSM film was prepared using the method reported in the reference [27]. Glass substrates (VWR Microscope slides, 2.5 cm by 2.5 cm) and ITO-coated glass substrates (1.2 cm by 0.7 cm, for conductive AFM measurement) were cleaned by four sequential ultrasonic baths in detergent, DI water, acetone and isopropyl alcohol. They were then further cleaned by oxygen plasma for 10 min. A series of pristine and hybrid PEDOT:PSS thin films with a thickness of about 40 nm was obtained by spin-coating. After spin-coating, all films were subsequently annealed inside a N2-filled glovebox at 230 °C for 30 min.

2.3 Measurement techniques

UV-Vis absorbance was performed using a Varian Cary 5E UV-Visible-NIR spectrometer. TEM characterizations were undertaken using a JEOL 2010 field-emission gun microscope operated at 200 kV. The surface profile and electrical conductivity were investigated by CSI Nano-Observer AFM with Resiscope II electrical module. Spectral absorptance was characterized by an integrating sphere (Labsphere RT-060). BRDF was investigated using the STIL-Reflet-90 Goniophotometer under white light from a halogen lamp. Sheet resistance was obtained from Cascade 4-point probe (CASCADE MICROTECH C4S-67 PROBE) together with a Keithley 238 source measurement unit and a Keithley 199 system digital multimeter scanner.

3. Results and discussion

3.1 Characterizations of Ag NPSMs

To obtain the basic optical properties of Ag NPSMs, a series of UV-Vis absorbance spectra were performed on as-synthesized Ag NPSM solutions (Fig. 1).Different batches of Ag NPSMs were obtained with the use of different amounts of seed solution during synthesis. While the thickness of Ag NPSMs is not sensitive to the amount of seed solution used, their final edge lengths increase as the amount of seed solution decreases. All spectra exhibit three characteristic absorbance peaks: The first two are located at 340 nm and 400 nm, corresponding respectively to the out-of-plane quadrupole and dipole LSPR bands. These two peaks remain unchanged when the edge-length of Ag NPSM increases. In addition, there is a main LSPR absorbance peak due to in-plane dipole resonance which is increasingly red-shifted (from 400 nm to 560 nm) as the aspect-ratio (between the edge-length and the thickness) increases. These optical properties are coherent with those published on similar Ag NPSMs [26,28]. The inset of Fig. 1 shows the TEM image of a few flat-lying Ag NPSMs and the solution absorbance of which shows an in-plane LSPR band at 560 nm. Despite some polydispersity on their dimensions, the majority of them show an edge-length of about 20 nm.

 figure: Fig. 1

Fig. 1 Normalized solution UV-Vis spectra of Ag NPSMs synthesized from different volumes of seed solution: black - seeds only; red - 2 mL; blue - 1mL; dark cyan - 650 μL; pink - 500 μL. The use of different amount of seed solution leads to different batches of Ag NPSMs of different edge-length/thickness aspect ratios. Inset: TEM image of flat-lying Ag NPSMs which have an in-plane LSPR band at 560 nm.

Download Full Size | PDF

Figure 2 shows the high resolution TEM images of a flat-lying Ag NPSM (left figure) and a few Ag NPSMs standing on their edges (right figure). A spacing of 0.235 nm can be measured in the right figure of Fig. 2 which corresponds to the periodicity of the {111} planes in FCC silver. Besides this lattice spacing, the thickness of the Ag NPSMs was measured to be abound 5 nm. The spacing of 0.250 nm measured from the left figure can corresponds to the periodicity observed along <111> directions and this spacing has been associated with the formation of stacking faults according to previous studies [26].

 figure: Fig. 2

Fig. 2 Left figure: High resolution TEM image of a flat-lying Ag NPSM. Right figure: High resolution TEM image of Ag NPSMs standing on their edges. Insets: the corresponding fast Fourier transform of the image.

Download Full Size | PDF

In brief summary, we synthesized a series of Ag NPSMs of similar thickness but of different edge length. By tuning the aspect-ratio between the edge length and the thickness, these different batches of Ag NPSMs exhibit a main in-plane LSPR peak at different wavelength in the visible spectrum.

3.2 Characterizations of hybrid PEDOT:PSS films

To study the optical properties of hybrid Ag NPSM-PEDOT:PSS thin films, their absorptance and Bidirectional Reflectance Distribution Functions (BRDF) were investigated. While absorbance was used to characterize the optical properties for solution samples, spectral absorptance (i.e. the direct proportional ratio of the radiation absorbed by the thin film to that incident upon it), was measured by an integrating sphere and used to characterize the optical properties of thin films.

3.2.1 Absorptance

We selected the batch of Ag NPSMs which exhibits a main solution LSPR band at 560 nm and we applied these NPSMs to fabricate hybrid Ag NPSM-PEDOT:PSS films. NPSM solutions of different concentration of 0.6 mg/mL, 1.6 mg/mL and 2.6 mg/mL were respectively used to formulate different hybrid films containing different amount of Ag NPSMs. Figure 3 shows the absorptance spectra of these hybrid films together with the absorptance of films containing PEDOT:PSS only and Ag NPSMs only. Compared to the solution absorbance spectra shown in Fig. 1, the thin film absorptance of Ag NPSMs exhibited significant LSPR band broadening and blue-shift which likely originates from their aggregation as well as the shortening of edge length due to PSS etching [29,30]. Compared to pristine PEDOT:PSS films, there are new absorption bumps in hybrid films in the wavelength range between 450 nm and 600 nm which are associated with the addition of Ag NPSMs. Besides these new absorption bumps, the overall absorptance of hybrid films increases with larger amount of Ag NPSMs used.

 figure: Fig. 3

Fig. 3 Absorption spectra of pure Ag NPSM film, pristine and hybrid PEDOT:PSS films of different Ag NPSM concentrations.

Download Full Size | PDF

3.2.2 Bidirectional Reflectance Distribution Functions (BRDF)

BRDF are defined by the CIE (Commission Internationale de l’Eclairage) as:

BRDF(θd,φd,θi,φi)=L(θd,φd)/E(θi,φi)

Where L(θd,φd) is the sample luminance in the direction (θd,φd), E(θi,φi) is the illuminance on the sample surface in the direction (θi,φi). The units of L(θd,φd), E(θi,φi) and BRDF are, respectively: candela/m2, lux, sr−1. BRDF can describe quantitatively the reflectance behavior of a non-lambertian surface by the variation of direction of view. In this work they are used to study the optical scattering properties of hybrid Ag NPSM-PEDOT:PSS thin films.

Figure 4 shows the difference of BRDF (ΔBRDF) between different hybrid Ag NPSM-PEDOT:PSS films and the pristine PEDOT:PSS film in both the in-plane (the incident light plane) and the out-of-plane directions. As shown in the in-plane ΔBRDF (left figure of Fig. 4), at 8° which is the specular reflection direction, the specular reflection dominates and the intensity increases with increasing Ag NPSM concentration. Except for the specular reflection feature, there is a remarkable resemblance between the in-plane and the out-of-plane BRDF for each hybrid film suggesting film uniformity. In low-angle regions such as (−14° to 3°) and (13° to 30°) for the in-plane configuration and (−20° to 20°) for the out-of-plane configuration, due to the constant amount of total incoming light, smaller ΔBRDF values were observed on samples where more light was reflected or scattered at other angles. For pure Ag NPSM films, Ag absorption dominates in this angular region while for hybrid films their ΔBRDF contain the compromise between Ag absorption and scattering. In high angle regions, such as (−90° to −14°) and (30° to 90°) for the in-plane configuration and (−90° to −20°) and (20° to 90°) for the out-of-plane configuration, the portion of light scattered by Ag NPSMs becomes larger and larger as the angle increases and this becomes the dominant feature of the spectra. Comparing hybrid films of different NPSM concentration, in these high angle regions the amount of scattered light increases significantly with increasing Ag NPSM concentration. Such increased light scattering is believed to be beneficial to harvest light in organic solar cells.

 figure: Fig. 4

Fig. 4 In-plane (in the incident light plane, named as in-plane) ΔBRDF (left) and out-of-plane (in the plane perpendicular to the incident plane) ΔBRDF (right) obtained from hybrid PEDOT:PSS films of various Ag NPSM concentrations.

Download Full Size | PDF

3.2.3 Surface profile and electrical conductivity

In order to study the surface morphology and electrical properties, we performed AFM, conductive AFM (CSI Nano-Observer AFM with Resiscope II electrical module) and four-point-probe resistance measurements on hybrid PEDOT:PSS films of different Ag NPSM concentration.

AFM height images (Fig. 5) show increased film roughness associated with the increased Ag NPSM concentration in PEDOT:PSS. On these images there are visible bright spots of various dimensions corresponding to individual Ag NPSM and agglomerates of multiple NPSMs. The formation of agglomerates is difficult to prevent in the method used to produce these hybrid films [3]. Such nanoparticle agglomeration leads to areas of larger heights compared to the average film thickness and increased surface roughness.

 figure: Fig. 5

Fig. 5 AFM height images (a, c, d, e) and C-AFM resistance signal images (b, d, f, h) for different Ag NPSM concentrations in PEDOT:PSS: (a, b) 0 mg/mL; (c, d) 0.6 mg/mL; (e, f) 1.6 mg/mL; and (g, h) 2.6 mg/mL. All images have a lateral scale of 2 by 2 microns.

Download Full Size | PDF

On the same sample region we also performed C-AFM characterizations (Fig. 5). For each hybrid film the bright spots appeared in AFM correspond well to regions of low local resistance observed in C-AFM. The observed local resistance reduction is believed to originate from the metallic nature of Ag NPSMs and their agglomerates or from regions where Ag NPSMs form chemical bonds with the sulfur atoms of PEDOT:PSS [20]. Such bonding formation can lead to a replacement of the insulating PSS by Ag NPSMs and thus reduce the local resistance. Except for these local low-resistance regions, as the concentration of NPSM increases, higher resistance is however observed in other regions possibly corresponding to the organic matrix. As a result, the overall film resistance becomes higher and higher as the amount of NPSM increases.

This is consistent with our sheet resistance measurements on these pristine and hybrid films by the 4-point-probe technique (Fig. 6): the sheet resistance of hybrid PEDOT:PSS films increases with increasing Ag NPSM concentration. This may originate from the insulating PSS polymers contained in Ag NPSM solutions since excess amount of PSS was used as stabilizers during the synthesis. Indeed, upon increasing the number of post-synthesis NPSM purification which involves particle precipitation and their re-dissolution in DI water to eliminate excess organic stabilizers, we observed a significant decrease of sheet resistance in the hybrid films for a fixed concentration of Ag NPSMs to a resistance value comparable to those found in pristine PEDOT:PSS films (Fig. 6).

 figure: Fig. 6

Fig. 6 Sheet resistance of hybrid Ag NPSM-PEDOT:PSS films as a function of Ag NPSM concentration (dark squares). The red dot shows the sheet resistance of hybrid Ag NPSM-PEDOT:PSS films after increasing the number of times of purification (involving particle precipitation by adding non-solvent THF and their re-dissolution in DI water). The blue triangle shows the sheet resistance of hybrid Ag NPSM-PEDOT:PSS films after combining both increased extent of purification and organic solvent additives.

Download Full Size | PDF

The possibility of reducing the global sheet resistance of hybrid Ag NPSM-PEDOT:PSS films through nanoparticle purification is important for their applications in organic solar cells because high sheet resistance of the hybrid layer can lead to high series resistance and poor fill-factor of the solar cell. Besides post-synthesis purification, it is also possible to use organic solvent additives [31,32] to reduce the sheet resistance of the hybrid films. In this second approach the addition of a small volume percentage of organic solvent additives such as dimethyl sulfoxide, ethylene glycol, and glycerol has been found highly effective in reducing the resistance of PEDOT:PSS films possibly through inducing a conformational change of the PEDOT chains [9,33]. In this work 5% (v/v) of glycerol was added into the hybrid Ag NPSM-PEDOT:PSS solution before fabricating hybrid thin films by spin-coating. The use of glycerol additive led to a dramatic reduction of sheet resistance (Fig. 6, blue triangle, corresponding to 0.06 MOhm/sq) to a resistance value even smaller than those observed in pristine PEDOT:PSS films. These results show that there are multiple approaches available, such as nanoparticle purification and organic solvent additives, to partially or completely remove the disadvantage of increased sheet resistance observed in hybrid Ag NPSM-PEDOT:PSS films compared to pristine PEDOT:PSS films.

4. Conclusion

The optical, structural, and electrical properties of hybrid PEDOT:PSS thin films doped with different amounts of Ag NPSMs have been investigated by various characterization techniques including transmission electron microscopy, regular and conductive atomic force microscopy, optical absorption, goniophotometry, and four-point probe resistance measurement.

The addition of Ag NPSMs has a strong effect on the optical properties of the PEDOT:PSS films: (1) Increased optical absorption was found as the amount of Ag NPSM increased; and (2) more light is scattered at high angle associated with the addition of Ag NPSMs. The increased light scattering in hybrid films can possibly lead to more efficient light harvest in organic solar cells.

The addition of Ag NPSMs also modified the structural and electrical properties of the PEDOT:PSS films: (1) Increased surface roughness was observed upon the incroporation of Ag NPSMs which is likely due to the formation of Ag agglomerates; and (2) global increased sheet resistance was found in hybrid films despite the formation of local low resistance regions from Ag NPSMs. While the increased surface roughness in hybrid films can be potentially improved by future advance in colloidal surface chemistry, we showed that the sheet resistance in hybrid films can be reduced to a resistance value comparable to or lower than those in pristine films by increasing the extent of NPSM post-synthesis purification and/or by applying organic solvent additives.

Acknowledgment

We would like to thank Dr. XU Xiangzhen (Laboratoire de Physique et d'Etude des Matériaux, ESPCI/CNRS/UPMC) for her help in electron microscopy characterizations, Concept Scientific Instruments and ScienTec for AFM and C-AFM characterizations. We acknowledge the China Scholarship Council-Ecole Centrale Marseille, CNRS 7334 & 8213 and IM2NP, Aix-Marseille University for financial support. Z. Chen acknowledges support from the ANR-2011-JS09-004-01-PvCoNano project and the EU Marie Curie Career Integration Grant (nº 303824).

References and links

1. S. Li, M. Meng Lin, M. S. Toprak, K. Kim, and M. Muhammed, “Nanocomposites of polymer and inorganic nanoparticles for optical and magnetic applications,” Nano Rev 1(0), 1–19 (2010). [CrossRef]   [PubMed]  

2. O. M. Folarin, E. R. Sadiku, and A. Maity, “Polymer-noble metal nanocomposites: Review,” Int. J. Phys. Sci. 6, 4869–4882 (2011).

3. G. Kickelbick, “Concepts for the incorporation of inorganic building blocks into organic polymers on a nanoscale,” Prog. Polym. Sci. 28(1), 83–114 (2003). [CrossRef]  

4. D. W. Hatchett and M. Josowicz, “Composites of intrinsically conducting polymers as sensing nanomaterials,” Chem. Rev. 108(2), 746–769 (2008). [CrossRef]   [PubMed]  

5. B. C. Sih and M. O. Wolf, “Metal nanoparticle-conjugated polymer nanocomposites,” Chem. Commun. (Camb.) (27): 3375–3384 (2005). [CrossRef]   [PubMed]  

6. S. Woo, J. H. Jeong, H. K. Lyu, Y. S. Han, and Y. Kim, “In situ-prepared composite materials of PEDOT: PSS buffer layer-metal nanoparticles and their application to organic solar cells,” Nanoscale Res. Lett. 7(1), 641 (2012). [CrossRef]   [PubMed]  

7. S. Pothukuchi, Y. Li, and C. P. Wong, “Development of a novel polymer-metal nanocomposite obtained through the route of in situ reduction for integral capacitor application,” J. Appl. Polym. Sci. 93(4), 1531–1538 (2004). [CrossRef]  

8. Z. Hu, J. Zhang, Z. Hao, and Y. Zhao, “Influence of doped PEDOT:PSS on the performance of polymer solar cells,” Sol. Energy Mater. Sol. Cells 95(10), 2763–2767 (2011). [CrossRef]  

9. A. Elschner, PEDOT Principles and Applications of an Intrinsically Conductive Polymer (Taylor and Francis Group, LLC, 2011).

10. H. Frohne, S. E. Shaheen, C. J. Brabec, D. C. Müller, N. S. Sariciftci, and K. Meerholz, “Influence of the anodic work function on the performance of organic solar cells,” ChemPhysChem 3(9), 795–799 (2002). [CrossRef]   [PubMed]  

11. Z. Su, L. Wang, Y. Li, H. Zhao, B. Chu, and W. Li, “Ultraviolet-ozone-treated PEDOT:PSS as anode buffer layer for organic solar cells,” Nanoscale Res. Lett. 7(1), 465 (2012). [CrossRef]   [PubMed]  

12. S.-W. Baek, J. Noh, C.-H. Lee, B. Kim, M.-K. Seo, and J.-Y. Lee, “Plasmonic forward scattering effect in organic solar cells: a powerful optical engineering method,” Sci. Rep. 3, 1726 (2013). [CrossRef]  

13. D. Duche, P. Torchio, L. Escoubas, F. Monestier, J.-J. Simon, F. Flory, and G. Mathian, “Improving light absorption in organic solar cells by plasmonic contribution,” Sol. Energy Mater. Sol. Cells 93(8), 1377–1382 (2009). [CrossRef]  

14. S. Vedraine, P. Torchio, D. Duché, F. Flory, J.-J. Simon, J. Le Rouzo, and L. Escoubas, “Intrinsic absorption of plasmonic structures for organic solar cells,” Sol. Energy Mater. Sol. Cells 95, S57–S64 (2011). [CrossRef]  

15. M. A. G. Namboothiry, T. Zimmerman, F. M. Coldren, J. Liu, K. Kim, and D. L. Carroll, “Electrochromic properties of conducting polymer metal nanoparticles composites,” Synth. Met. 157(13-15), 580–584 (2007). [CrossRef]  

16. H. A. Atwater and A. Polman, “Plasmonics for improved photovoltaic devices,” Nat. Mater. 9(3), 205–213 (2010). [CrossRef]   [PubMed]  

17. S. Pillai and M. A. Green, “Plasmonics for photovoltaic applications,” Sol. Energy Mater. Sol. Cells 94(9), 1481–1486 (2010). [CrossRef]  

18. K. J. Moreno, I. Moggio, E. Arias, I. Llarena, S. E. Moya, R. F. Ziolo, and H. Barrientos, “Silver nanoparticles functionalized in situ with the conjugated polymer (PEDOT:PSS),” J. Nanosci. Nanotechnol. 9(6), 3987–3992 (2009). [CrossRef]   [PubMed]  

19. R. G. Melendez, K. J. Moreno, I. Moggio, E. Arias, A. Ponce, I. Llanera, and S. E. Moya, “On the Influence of Silver Nanoparticles Size in the Electrical Conductivity of PEDOT: PSS,” Mater. Sci. Forum 644, 85–90 (2010). [CrossRef]  

20. N. G. Semaltianos, W. Perrie, S. Romani, R. J. Potter, G. Dearden, and K. G. Watkins, “Polymer-nanoparticle composites composed of PEDOT:PSS and nanoparticles of Ag synthesised by laser ablation,” Colloid Polym. Sci. 290(3), 213–220 (2012). [CrossRef]  

21. H. S. Noh, E. H. Cho, H. M. Kim, Y. D. Han, and J. Joo, “Organic solar cells using plasmonics of Ag nanoprisms,” Org. Electron. 14(1), 278–285 (2013). [CrossRef]  

22. L. Lu, Z. Luo, T. Xu, and L. Yu, “Cooperative plasmonic effect of Ag and Au nanoparticles on enhancing performance of polymer solar cells,” Nano Lett. 13(1), 59–64 (2013). [CrossRef]   [PubMed]  

23. J.-L. Wu, F.-C. Chen, Y.-S. Hsiao, F.-C. Chien, P. Chen, C.-H. Kuo, M. H. Huang, and C.-S. Hsu, “Surface plasmonic effects of metallic nanoparticles on the performance of polymer bulk heterojunction solar cells,” ACS Nano 5(2), 959–967 (2011). [CrossRef]   [PubMed]  

24. V. Bastys, I. Pastoriza-Santos, B. Rodríguez-González, R. Vaisnoras, and L. M. Liz-Marzán, “Formation of Silver Nanoprisms with Surface Plasmons at Communication Wavelengths,” Adv. Funct. Mater. 16(6), 766–773 (2006). [CrossRef]  

25. F. Nicodemus, J. Richmond, J. Hsia, I. Ginsburg, and T. Limperis, Geometrical Considerations and Nomenclature for Reflectance (U.S. Dept. of Commerce, National Bureau of Standards: for sale by the Supt. of Docs, U.S. Govt. Print. Off, 1977).

26. D. Aherne, D. M. Ledwith, M. Gara, and J. M. Kelly, “Optical properties and growth aspects of silver nanoprisms produced by a highly reproducible and rapid synthesis at room temperature,” Adv. Funct. Mater. 18(14), 2005–2016 (2008). [CrossRef]  

27. A. P. Kulkarni, K. M. Noone, K. Munechika, S. R. Guyer, and D. S. Ginger, “Plasmon-enhanced charge carrier generation in organic photovoltaic films using silver nanoprisms,” Nano Lett. 10(4), 1501–1505 (2010). [CrossRef]   [PubMed]  

28. G. S. Metraux and C. A. Mirkin, “Rapid thermal synthesis of silver nanoprisms with chemically tailorable thickness,” Adv. Mater. 17, 412–415 (2005). [CrossRef]  

29. D. Aherne, D. E. Charles, M. E. Brennan-Fournet, J. M. Kelly, and Y. K. Gun’ko, “Etching-resistant silver nanoprisms by epitaxial deposition of a protecting layer of gold at the edges,” Langmuir 25(17), 10165–10173 (2009). [CrossRef]   [PubMed]  

30. T. Pal, T. K. Sau, and N. R. Jana, “Reversible formation and dissolution of silver nanoparticles in aqueous surfactant media,” 7463, 1481–1485 (1997).

31. J. Y. Kim, J. H. Jung, D. E. Lee, and J. Joo, “Enhancement of electrical conductivity of poly(3,4-ethylenedioxythiophene)/poly(4-styrenesulfonate) by a change of solvents,” Synth. Met. 126(2-3), 311–316 (2002). [CrossRef]  

32. O. P. Dimitriev, D. A. Grinko, Y. V. Noskov, N. A. Ogurtsov, and Pud, “PEDOT:PSS films-Effect of organic solvent additives and annealing on the film conductivity,” Synth. Met. 159(21-22), 2237–2239 (2009). [CrossRef]  

33. J. Ouyang, Q. Xu, C.-W. Chu, Y. Yang, G. Li, and J. Shinar, “On the mechanism of conductivity enhancement in poly(3,4-ethylenedioxythiophene):poly(styrene sulfonate) film through solvent treatment,” Polymer (Guildf.) 45(25), 8443–8450 (2004). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (6)

Fig. 1
Fig. 1 Normalized solution UV-Vis spectra of Ag NPSMs synthesized from different volumes of seed solution: black - seeds only; red - 2 mL; blue - 1mL; dark cyan - 650 μL; pink - 500 μL. The use of different amount of seed solution leads to different batches of Ag NPSMs of different edge-length/thickness aspect ratios. Inset: TEM image of flat-lying Ag NPSMs which have an in-plane LSPR band at 560 nm.
Fig. 2
Fig. 2 Left figure: High resolution TEM image of a flat-lying Ag NPSM. Right figure: High resolution TEM image of Ag NPSMs standing on their edges. Insets: the corresponding fast Fourier transform of the image.
Fig. 3
Fig. 3 Absorption spectra of pure Ag NPSM film, pristine and hybrid PEDOT:PSS films of different Ag NPSM concentrations.
Fig. 4
Fig. 4 In-plane (in the incident light plane, named as in-plane) ΔBRDF (left) and out-of-plane (in the plane perpendicular to the incident plane) ΔBRDF (right) obtained from hybrid PEDOT:PSS films of various Ag NPSM concentrations.
Fig. 5
Fig. 5 AFM height images (a, c, d, e) and C-AFM resistance signal images (b, d, f, h) for different Ag NPSM concentrations in PEDOT:PSS: (a, b) 0 mg/mL; (c, d) 0.6 mg/mL; (e, f) 1.6 mg/mL; and (g, h) 2.6 mg/mL. All images have a lateral scale of 2 by 2 microns.
Fig. 6
Fig. 6 Sheet resistance of hybrid Ag NPSM-PEDOT:PSS films as a function of Ag NPSM concentration (dark squares). The red dot shows the sheet resistance of hybrid Ag NPSM-PEDOT:PSS films after increasing the number of times of purification (involving particle precipitation by adding non-solvent THF and their re-dissolution in DI water). The blue triangle shows the sheet resistance of hybrid Ag NPSM-PEDOT:PSS films after combining both increased extent of purification and organic solvent additives.

Equations (1)

Equations on this page are rendered with MathJax. Learn more.

BRDF( θ d , φ d , θ i , φ i )=L( θ d , φ d )/E( θ i , φ i )
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.