GLORIA

GEOMAR Library Ocean Research Information Access

Your email was sent successfully. Check your inbox.

An error occurred while sending the email. Please try again.

Proceed reservation?

Export
  • 1
    In: Japanese Journal of Applied Physics, IOP Publishing, Vol. 59, No. SD ( 2020-03-01), p. SDDC04-
    Type of Medium: Online Resource
    ISSN: 0021-4922 , 1347-4065
    RVK:
    RVK:
    RVK:
    Language: Unknown
    Publisher: IOP Publishing
    Publication Date: 2020
    detail.hit.zdb_id: 218223-3
    detail.hit.zdb_id: 797294-5
    detail.hit.zdb_id: 2006801-3
    detail.hit.zdb_id: 797295-7
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 2
    In: ECS Meeting Abstracts, The Electrochemical Society, Vol. MA2018-03, No. 4 ( 2018-07-13), p. 238-238
    Abstract: Organic photovoltaics (OPVs) cells are spotlighted as a one of the candidates for future energy source owing to their flexibility, low-cost and high-performance [ref. 1]. In a relatively short period, the power conversions (PCEs) of OPVs have achieved over 11% so that a variety of the recent researches such as improvement in molecular design and synthesis of new polymers and in improved solar cell device design has been successfully conducted to improve the OPVs performances. However, PCEs of other kinds of solar cells such as silicon solar cell and compound semiconductor solar cell have already reached over 20%. The reason why PCEs of OPVs is lower than one of the others is a recombination of carriers by Coulomb force, which is one of the biggest issues in OPVs. In many of the most efficient polymer-fullerene OPVs devices, 50% or more of the energy loss is caused by the recombination of carriers [ref. 2]. Recently, several reports to solve this problem have been reported. One of the promising ways is incorporating Ferroelectric co-polymer poly(vinylidenefluoride-trifluoroethylene) (P(VDF-TrFE)) between an active layer and electrode interface as a Langmuir-Blodgett films in OPV [ref. 3]. Ferroelectric polymers show electric field derived from its spontaneous polarization due to difference in electronegativity of fluorine and hydrogen. It is therefore possible to enhance in charge separation efficiency and carrier mobility of electron for suppression of recombination. In this research, we have successfully prepared P(VDF-TrFE) nanocrystals as s dispersion state by the reprecipitation method so that we could incorporate the nanocrystals in active layer in OPV [ref. 4]. P(VDF-TrFE) nanocrystals prepared by this method are approximately 40 nm. The obtained dispersion was mixed to dissolve with photoactive materials (P3HT : PCBM) at ratio of 15.0 : 10.5 by weight instead of a pure solvent. Moreover, the mixed solution was applied for OPV cells as an active layer. Short-circuit current (Jsc) and PCEs of OPV incorporating P(VDF-TrFE) nanocrystals was improved in at most 20% and 40%, respectively comparing to the one without P(VDF-TrFE) nanocrystals. The electric field derived from P(VDF-TrFE) nanocrystals and carrier mobility were estimated from J-V curve and calculation, respectively. Incident photo to current conversion efficiency (IPCE) of OPVs cells which are incorporating P(VDF-TrFE) was three times as high as the one which was not incorporating P(VDF-TrFE) although they had almost same absorbance and peak positions. Therefore, it was confirmed that P(VDF-TrFE) improved in charge separation efficiency and carrier mobility of electron, but not influenced light absorption. The details in experimental conditions and characteristics of P(VDF-TrFE) nanocrystals and OPVs, morphology and so on will be discussed in this presentation. References: [1] Erik J. et al., ACS Nano , 6, 4708-4714 (2013). [2] Yongbo Yuan. et al., Nat. Mater., 10, 296-302 (2011). [3] Zhengguo X. et al., Adv. Energy Mater ., 3, 1581-1588 (2013). [4] Nakanishi H. et al., in Single Organic Nanoparticles , eds. Masuhara H. et al., (Springer, Berlin, 2003) Chap. 2.
    Type of Medium: Online Resource
    ISSN: 2151-2043
    Language: Unknown
    Publisher: The Electrochemical Society
    Publication Date: 2018
    detail.hit.zdb_id: 2438749-6
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 3
    In: ECS Meeting Abstracts, The Electrochemical Society, Vol. MA2018-03, No. 4 ( 2018-07-13), p. 242-242
    Abstract: [Introduction] Organic charge transfer crystals (CTCs) can be interesting candidates as light absorbers of solar cells to eliminate large voltage loss inherent in dye-sensitized and bulk hetero junction types, in which carrier generation relies on the energy cascades. We have synthesized novel CTCs by forming salts between 1,3-bis(dicyanomethylidene)indan anion (TCNIH - ) as a donor and viologen cations as acceptors, namely, N,N’ -alkyl substituted 4,4’-bipyridiniums (Fig. 1, methyl = MV 2+ , ethyl = EV 2+ , heptyl = HV 2+ and octyl = OV 2+ ). [Experimental] Mixed salts of TCNIH - and viologens were obtained by slowly evaporating solvent at room temperature from their 2 : 1 mixed solution in ethanol. While their crystal structures were examined on powder samples and single crystals, their optical properties were studied by measuring UV-Vis and PL spectra between 77 and 298 K on a Horiba Fluorolog-3 equipped with a liq. N 2 cryostat. [Results and Discussion]   Colorless TCNIH 2 turns into deep blue in ethanol due to deprotonation from its methylene carbon to become TCNIH - anion (Fig. 1). While its salt with Na + was purplish black, having similar absorption features as the solution of TCNIH - with its onset around 800 nm, co-crystals with MV 2+ and EV 2+ were black with metallic shine to exhibit extended CT absorption in NIR up to ca. 1,000 nm (Fig. 2). On the other hand, those with HV 2+ and OV 2+ are quite similar to the Na + salt with only small extension. All of the mixed crystals exhibited XRD patterns different from those of original TCNIH 2 and viologen halides, indicating formation of their salts. Single crystal XRD structural analysis performed on large grains revealed monoclinic structure for MV 2+ salt, while triclinic for the rest, in which TCNIH - and viologen cations are alternately stacked in 2 : 1 ratio. On increasing the length of the alkyl substituents, expansion and distortion of lattice occurs as seen from their lattice constants summarized in Table 1. The close packing of TCNIH - and MV 2+ /EV 2+ by coulombic interaction obviously contributed to the enhancement of CT, whereas bulky HV 2+ and OV 2+ prohibited their intimate interaction. PL spectra were measured for the salts (Fig. 3). The Na + salt indicates a spectrum very similar to that of the TCNIH - solution with not much enhancement and peak shift between 77 and 298 K, indicating emission from the exciton localized in a single TCNIH - . The salts with MV 2+ and EV 2+ exhibit broad emissions in the NIR range, which are greatly enhanced and resolved into two peaks at 77 K. Exciton-phonon coupling to delocalize exciton within CTCs explains these observations. PL spectra of HV 2+ and OV 2+ salts are unique, with their energy clearly smaller than that of the Na + salt, despite of their very similar energy of excitation. Also, clear enhancement and resolution into two peaks at 77 K in case of OV 2+ salt suggest relaxation of the TCNIH - localized exciton down to the CT states within these crystals. Acknowledgment The present work was supported by Program for Advancing Strategic International Networks to Accelerate the Circulation of Talented Researchers, “Advanced Next Generation Energy Leadership (R2601, FY2014-2016)” of Japan Society for the Promotion of Science (JSPS). Figure 1
    Type of Medium: Online Resource
    ISSN: 2151-2043
    Language: Unknown
    Publisher: The Electrochemical Society
    Publication Date: 2018
    detail.hit.zdb_id: 2438749-6
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 4
    In: ECS Meeting Abstracts, The Electrochemical Society, Vol. MA2020-01, No. 16 ( 2020-05-01), p. 1095-1095
    Abstract: Combination of ionic organic donor and acceptor to form crystalline salt can be an interesting strategy to obtain new types of organic semiconductors, since charge transfer (CT) excitonic properties are anticipated. We have previously achieved a novel CT salt from deprotonated 1,3-(bisdicyanomethylidene)indan anion (TCNIH - ) and methyl viologen cation (MV 2+ ) in a 2:1 composition. In this study, the n-alkyl chain length in viologen was systematically changed to find out the border for the evolution of CT character of the mixed salts. [1] Propyl to hexyl viologens (PrV to HxV) were synthesized by quaternization of 4,4’-bipyridine using the corresponding 1-haloalkanes in the yields of 60-62%, whereas methyl, ethyl, heptyl and octyl viologens (MV, EV, HpV, OV) were commercially available. Co-crystals with TCNIH - were precipitated by mixing with viologens at 2 : 1 ratio in water, filtrated and recrystallized from ethanolic solutions. While their crystal structures were examined on powder samples and single crystals, their optical properties were studied by measuring UV-vis and photoluminescence (PL) spectra for CT excitation. Co-crystals with TCNIH - were obtained for all viologens in a 2:1 ratio in the yields of 39-44%. While Na + salt with TCNIH - preserves the character of the deprotonated TCNIH - in solution, showing a peak at around 590 nm and its absorption extending up to around 850 nm, those with RV 2+ having short alkyl chains exhibit absorption towards NIR range and concomitant decrease of visible absorption, due to CT from TCNIH - to RV 2+ (Fig. 1). A clear border can be found between PnV and HxV, as the CT absorption is significantly decreased when R is longer than Hx. Crystal structures of the mixed salts were identified as monoclinic for MV, BV, HxV salts, while triclinic for the others. The powder XRD data as well as melting points determined by differential scanning calorimetry indicated significant enlargement of donor-acceptor distances, decrease of melting point, and thereby weakened interaction to account for the observed CT characters. Coulombic interaction between donor and acceptor result in close packing of these co-crystals for their CT characters, whereas it is hindered when alkyl chains longer than Hx is introduced. However, PL from CT states was observed for all the compounds, as the salts with HxV, HpV and OV also showed clearly longer emission than that of Na salt, indicating primary excitation confined in TCNIH - , followed by its relaxation down to the CT states in such salts. Reference [1] E. Saito, et al., ECS Trans. , 2018 , 88 , 301-311 Figure 1
    Type of Medium: Online Resource
    ISSN: 2151-2043
    Language: Unknown
    Publisher: The Electrochemical Society
    Publication Date: 2020
    detail.hit.zdb_id: 2438749-6
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 5
    In: ECS Meeting Abstracts, The Electrochemical Society, Vol. MA2018-03, No. 4 ( 2018-07-13), p. 243-243
    Abstract: Charge-transfer (CT) complex crystals composed of organic electron donor (D) and acceptor (A) have been extensively studied because of their attractive properties such as metallic conductivity, photoconductivity, ambipolar charge transport, and high career mobility. Numerous research efforts have been made for their use in devices such as organic field effect transistors and organic photocells. However, bulky single crystals of CT complexes are difficult to handle, specifically to be made into a form of a thin film needed for solar cell applications.  Nano/micro crystallization is a promising path to overcome the above mentioned problem. However, the growth of high-quality nanocrystals is imperative for photoconductivity but often hampered, since CT complexes are poorly soluble and vapor pressures of donor and acceptor molecules are imbalanced[1-2].  Here we presente the “reprecipitation method” as facile synthesis route to obtain high quality organic CT complex nanocrystals. The process of nanocrystallization can be explained in four steps[3]; 1) A small amount of a solution of the target compound in its good solvent is injected into a large amount of its poor solvent under vigorous stirring. 2) The tiny droplets formed by the injection diffuse into the poor solvent as the two solvents are chosen to be miscible. 3) Nucleation and crystal growth of the target compound proceed due to its supersaturation at the boundary of good/poor solvents. 4) As the good solvent is completely diffused, nanocrystals are stably dispersed in the poor solvents. All these events are expected to occur in a short time (within tens of milliseconds[4] ) and thus nanocrystals with kinetically stable structures are often obtained, which are different from thermodynamically stable single bulk crystal structures[5].  In this work, we have successfully applied the reprecipitation method to dibenzotetrathiafulvalene (DBTTF)-tetracyanoquinodimethane (TCNQ) and tetrathiafulvalene (TTF)-TCNQ to obtain its nanocrystal. The DBTTF-TCNQ nanocrystals were prepared by typical reprecipitation method from D-A mixed solution. The nanocrystals prepared in such a way have been found to possess kinetically stable crystal structure, different from the bulk material. Moreover, a direct current device employing a polycrystalline thin film of the nanocrystals prepared by simple filtration an equivalent photosensitivity (354.57 μA/W) with that of a bulk crystal (417.14 μA/W) despite of the presence of many grain boundaries, indicating a favorable molecular stacking in the nanocrystals for career transport. In addition, the TTF-TCNQ nanocrystals were prepared by two-step reprecipitation method. TTF (TCNQ) solution was injected to TCNQ (TTF) nanoparticle water dispersion. The TTF-TCNQ nanocrystals were formed when the nanoparticles reacted and the dispersion color was changed immediately from yellow to black.  Details of morphology, characteristic and crystallization process of the nanocrystals will be discussed in presentation. Reference [1] M. Hiraoka, T. Hasegawa, T. Yamada, Y. Takahashi, S. Horiuchi, and Y. Tokura, Adv. Mater. 19 , 3248 (2007). [2] Y. Takahashi, T. Hasegawa, Y. Abe, Y. Tokura, and G. Saito, Appl. Phys. Lett. 88 , 073504 (2006). [3] K. Shito, N. Ito, and A. Masuhara, Jpn. J. Appl. Phys. 54 , 06FK05-1 (2015). [4] J. Mori, Y. Miyashita, D Oliveira, H. Kasai, H. Oikawa, and H. Nakanishi, J. Cryst. Growth 311 , 553 (2009). [5] L. Huang, Q. Liao, H. Fu, J. Ma, and J. Yao, J. Mater. Chem. 20 , 159 (2010).
    Type of Medium: Online Resource
    ISSN: 2151-2043
    Language: Unknown
    Publisher: The Electrochemical Society
    Publication Date: 2018
    detail.hit.zdb_id: 2438749-6
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 6
    Online Resource
    Online Resource
    The Electrochemical Society ; 2018
    In:  ECS Meeting Abstracts Vol. MA2018-03, No. 4 ( 2018-07-13), p. 247-247
    In: ECS Meeting Abstracts, The Electrochemical Society, Vol. MA2018-03, No. 4 ( 2018-07-13), p. 247-247
    Abstract: Introduction: Combination of inorganic and organic materials offers unlimited new opportunities. While “composites” stands for their physical mixtures, “hybrids” and “inorganic/organic compounds” are especially interesting for their new functionalities, as they are achieved by self-organization due to “chemistry” among the constituents. We have established electrochemical self-assembly (ESA) of inorganic/organic hybrid thin films from one pot containing all the ingredients. 1,2) In this work, we have employed cathodic electrodeposition of CuSCN as the test bed for the ESA with various organic dyes to understand its principle and evaluated optical properties of the product thin films. Experiments: CuSCN / dye hybrid thin films were cathodically electrodeposited according to the reaction, [Cu(II)(SCN)] + + e - → Cu(I)SCN, in the presence of various organic dyes shown in Fig. 1. Photoluminescence (PL) and PL excitation (PLE) spectra were measured between 77 and 298 K on a Horiba Fluorolog-3 equipped with a high power Xe lamp excitation source and lq. N 2 -cooled PMT detector. Results and discussion: Anionic xanthene dyes which afford ESA with ZnO were not successful with CuSCN except FLNCS, whereas zwitter ionic RB and all the cationic dyes yielded colorful thin films as they were loaded into CuSCN. Soft and hard, acid and base (HSAB) principle nicely explains the behavior of FLNCS, as its –NCS is soft Lewis basic, finding stable coordination to soft Lewis acidic Cu(I) site of CuSCN, just like the hard –COO - being good with hard Zn(II) of ZnO. On the other hand, partial replacement of Cu + ions with cationic dyes (RB with its ammonium moiety) can account for their successful ESA. These two mechanisms were confirmed by the decrease of dye loading in baths containing excess SCN − and Cu 2+ , respectively. RB, R6G, NB and DAST show strong PL both as their solutions and in solid states with their PLEs nicely matched with their absorption spectra, while that of the hybrid thin films was totally quenched except for DAST (CuSCN/DAS hybrid thin film). The energy diagram drawn for the energy levels of individual components suggest hole transfer from HOMO of photoexcited organic chromophore to valence band of inorganic CuSCN, except for NB. Hybridization of NB with CuSCN might have altered its energetic structure to exhibit photo-induced carrier generation of the hybrid thin films, thus suggesting usefulness of the hybrid thin films with RB, R6G and NB as light absorbers in solar cells. On the other hand, the DAS + chromophore exhibits PL even when it is loaded into CuSCN, with its energy in between solution (= monomer) and powder of DAST. While PL of solution and powder become sharp, blue-shifted and intensified at 77 K, as typically expected by exciton confinement, that of the hybrid film is red-shifted and not much intensified, suggesting a strong excitonic interaction with CuSCN to offer a new type of luminescent material. References: T. Iwamoto et al., J. Phys. Chem. C, 118, 16581-16590 (2014). Y. Tsuda et al., Microsys. Tech., DOI 10.1007/s00542-017-3394-9. Figure 1
    Type of Medium: Online Resource
    ISSN: 2151-2043
    Language: Unknown
    Publisher: The Electrochemical Society
    Publication Date: 2018
    detail.hit.zdb_id: 2438749-6
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 7
    Online Resource
    Online Resource
    The Electrochemical Society ; 2018
    In:  ECS Meeting Abstracts Vol. MA2018-03, No. 4 ( 2018-07-13), p. 244-244
    In: ECS Meeting Abstracts, The Electrochemical Society, Vol. MA2018-03, No. 4 ( 2018-07-13), p. 244-244
    Abstract: 「 Introduction 」 Organic-inorganic perovskite crystals have been aggressively applied to perovskite solar cells due to their attractive properties such as long carrier diffusion, am-bipolar conductivity, broad color-tunability, and their small exciton binding energy 1 . These perovskite crystals have been also applied to various applications like light-emitting diodes (LED), photodetectors, field effect transistors, and lasers due to perovskite crystals can be simply prepared from low-cost precursors with simple solution processes 1 . These tremendous potential for various applications with high performance are based on perovskite electro-optical features and they are expected to be used as an alternative semiconductor material to silicon and compound materials. Recently, bright luminescence properties from methylammonium lead tri-bromide (MAPbBr 3 ) perovskite quantum dots (PeQDs) have been reported owing to the development of methods for preparing these PeQDs. Zhang et al. developed a ligand-assisted reprecipitation (LARP) inspired by the reprecipitation method for preparing organic and nano / micro crystals 2 . LARP can simply obtain MAPbX 3 (X = I, Br, and Cl) PeQDs through mixing a solution dissolving precursors and poor solvent. Although synthesis methods representative for LARP with narrow size-distributions are of particular interest for the successful implementation of PeQDs into LED with narrow emission, size-controlled PeQDs by nanometers have not been developed. In this work, we report on Ostwald ripening as a size-tunable technique for MAPbBr 3 PeQDs using LARP. 「 Experimental 」 Size-controlled synthesis of MAPbBr 3 PeQDs, and a simple strategy is as follows; a precursor solution was dropped into poor solvent followed by centrifuged to remove large-sized MAPbBr 3 crystals. Herein, MAPbBr 3 PeQD dispersions were aged at the 50℃using water bath to accelerate Ostwald ripening in MAPbBr 3 PeQD dispersions. Finally, we obtained size-controlled MAPbBr 3 PeQDs through centrifugation to remove grown MAPbBr 3 crystals. 「 Results and Discussion 」 MAPbBr 3 PeQD dispersions can be clearly observed that as aging time increases, PL peaks are blue shifted from 517 nm to 456 nm through aging for 4 hours. The emission colors irradiated UV lamp at 254 nm of MAPbBr 3 PeQD dispersions also changed from green to cyan and deep blue , while these dispersions close to be color less. When MAPbBr 3 PeQD dispersion is aged 1 hour, compared with an initial MAPbBr 3 PeQD dispersion, maximum PL peak was not clearly blue shift (ΔE = 42 meV) because of mean QD sizes were not drastically changed and also these square shape is maintained. In contrast, MAPbBr 3 PeQD dispersion is aged for 2 hours, a maximum PL peak drastically blue shifted from 508 nm to 489 nm. Because morphologies of samples such as sizes and shapes were completely changed from square to spherical with 7.2 nm size , result in PL energy extremely change (ΔE = 95 meV). Moreover, as further aging times increases, their mean QD sizes down to approach 5.3 nm by aging at the 50℃for 4 hours, in which clearly appear increasing band gap of MAPbBr 3 PeQDs due to quantum confinement effect. 「 References 」 1) B. R. Sutherland, et al., Nature , 2016 , 10, 295-302. 2) F. Zhang, et al., ACS Nano , 2015, 9, 4533-4542.
    Type of Medium: Online Resource
    ISSN: 2151-2043
    Language: Unknown
    Publisher: The Electrochemical Society
    Publication Date: 2018
    detail.hit.zdb_id: 2438749-6
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 8
    In: ECS Meeting Abstracts, The Electrochemical Society, Vol. MA2018-03, No. 4 ( 2018-07-13), p. 246-246
    Abstract: We have found electrochemical self-assembly (ESA) of inorganic / organic hybrid thin films in which the inorganic is CuSCN, known to be a wide bandgap p-type semiconductor, whereas the organic is 4- N , N -dimethylamino-4’- N ’-methylstilbazolium chromophore (abbreviated as DAS + ) as its salt with tosylate (DAST) is known to exhibit second-order nonlinear optical property for terahertz emitters [1]. DAS + is also known to yield a layered inorganic-organic hybrid crystal in a (DAS)(Cu 5 I 6 ) composition [2]. Strong dipole-dipole interaction of the DAS + chromophores aligned between the CuI layers results in a spontaneous photocarrier generation and ambipolar transport in a single absorber solar cell. Thus, CuSCN/DAS hybrid thin films can be attractive alternatives for such opto-electrical applications, when ordered arrangement of DAS+ chromophore is achieved during ESA. In our previous study, switching of dye loading mechanism has been suggested, depending on DAS + concentration in the bath [1]. In low DAS + concentration range, the loading is limited by diffusion so that DAS + is entrapped within CuSCN crystal grains, while surface reaction of hybridization begins to limit the dye loading in high DAS + concentration range, resulting in formation of unique nanostructures as well as phase separation of inorganic and organic domains. In this study, electrochemical analysis by employing rotating disk electrode (RDE) has been performed to verify the mechanism of ESA and also to explore the limit of DAS + loading. Electrodeposition of CuSCN undergoes as limited by transport of 1 : 1 complex between Cu 2+ and SCN - ions, namely {[Cu(SCN)] + } species, near 100% Faradic efficiency with marginal influence by the presence of DAS + . Thus, the rate of CuSCN growth is always proportional to the concentration of the active species, and also to ω 1/2 (ω = angular speed of rotation of RDE). On the other hand, the rate of DAS + precipitation, v(DAS)(SCN), should also be proportional to ω 1/2 under the regime of diffusion limited loading with a given [DAS + ], and is independent of CuSCN formation rate. When the rate surface complex formation begins to limit the rate of DAS + precipitation, the amount of DAS + loading should become proportional to the rate of CuSCN deposition, if a second order rate law as shown below holds, v(DAS)(SCN) (mol s -1 cm -2 ) = k[DAS + ][site] (1) where, k (mol -1 cm 3 s -1 ) is the second order reaction rate constant for formation of surface complex and [site] is the surface concentration of newly formed CuSCN sites, which should be expressed as, [site] (mol cm -2 ) = A × 0.62 × {[Cu(SCN)] + } × D{[Cu(SCN)] + } 2/3 × ν(methanol, 298 K) -1/6 × ω 1/2 (2) which is simply derived from Levich equation for diffusion limited electrodeposition of CuSCN, multiplied by a proportionality constant, A (s), to express the activity of the surface site. The amount of DAS + loaded into the film electrodeposited for a given [DAS + ] under variation of {[Cu(SCN)] + } in the bath was examined, and indeed such a trend as predicted from the above-mentioned model was found (Fig. 1). In the high {[Cu(SCN)] + } range, DAS + loading is independent of {[Cu(SCN)] + } and appears proportional to [DAS + ], because of the diffusion limited loading mechanism. When {[Cu(SCN)] + } goes below certain concentration, DAS + loading changes proportionally to {[Cu(SCN)] + }, for a given [DAS + ], but is also proportional to [DAS + ] for a given {[Cu(SCN)] + }, as expected for the surface reaction limitation expressed by Eq. (1). The switching from surface reaction to diffusion limitation occurs at the lower {[Cu(SCN)] + } when [DAS + ] goes lower, as recognized by the shift of the border between {[Cu(SCN)] + } dependent and independent parts. [1] Yuki Tsuda et al., Monatshefte für Chemie, 148, 845-854 (2017) [2] Elena Cariati et al., Adv. Mater., 13, 1665-1668 (2001). Figure 1
    Type of Medium: Online Resource
    ISSN: 2151-2043
    Language: Unknown
    Publisher: The Electrochemical Society
    Publication Date: 2018
    detail.hit.zdb_id: 2438749-6
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 9
    Online Resource
    Online Resource
    The Electrochemical Society ; 2018
    In:  ECS Meeting Abstracts Vol. MA2018-03, No. 4 ( 2018-07-13), p. 265-265
    In: ECS Meeting Abstracts, The Electrochemical Society, Vol. MA2018-03, No. 4 ( 2018-07-13), p. 265-265
    Abstract: Metal organic frameworks (MOFs) are widely studied because of their unique structures and wide applications represented by gas separation, gas sensor and electrocatalysis. In our previous research, 4 types of Zn-telephthalic acid (TPA) MOFs were synthesized via microwave hydrothermal reaction by varying pH of precursor solution. One of the Zn-TPA MOFs in a Zn 4 (OH) 6 (C 8 H 6 O 4 ) composition exhibited a stable reversible redox reaction accompanied with proton exchange owing to its structural rigidity brought by para positional carboxyl groups (–COOH), to show its usefulness as a negative electrode material for redox batteries. In this study, we have employed trimellitic acid (1,2,4-benzenetricarboxylic acid, abbreviated as TMLA) for synthesis of Zn-based MOFs. TMLA has an extra carboxyl groups to the structure of TPA, so that an increased storage capacity is anticipated. Aqueous solutions containing 0.1 M Zn (CH 3 COO) 2 and 0.035 M TMLA with pH adjusted between 7 and 5 by KOH were put into a Teflon-lined vessels for microwave (2.45 GHz) reaction at 150℃ for 30 min. The powder samples were centrifugally collected, washed and dried for analysis by XRD and TG-DTA. Paste containing 35wt% powder was coated F-doped tin oxide (FTO) glass to fabricate mesoporous electrodes with ca. 10 μm thickness for their evaluation by cyclic voltammetry and chronoamperometry in a 0.1 M KCl. Three types of Zn-TMLA MOFs were synthesized depending on pH and two of them in layered structures with 14.4 and 12.2 Å spacing could be isolated, with their compositions determined as Zn 5 (OH) 7 (C 9 H 3 O 6 ) and Zn 4 (OH) 5 (C 9 H 3 O 6 )•1.5H 2 O from TG-DTA, named as types A and B, respectively. Both A and B exhibit reversible redox reactions as shown in Fig. 1, with their coulombic reversibility of 85 and 90%, E 1/2 values of -1.12 and -1.10 V vs. Ag/AgCl, respectively (determined from 70 mV s -1 scans), similar to 90% and -1.03 V found for Zn-TPA MOF. Oxidation peaks are positively shifted on the slow end of the potential scanning, unlike the case with Zn-TPA, which might be associated with reconstruction of the MOF structure, upon proton exchange in the redox reactions expressed as follows, Zn 5 (OH) 7 (C 9 H 3 O 6 ) + 3H + + 3e - ⇄ Zn 5 (OH) 7 (C 9 H 6 O 6 ) (A) Zn 4 (OH) 5 (C 9 H 3 O 6 ) + 3H + + 3e - ⇄ Zn 4 (OH) 5 (C 9 H 6 O 6 ) (B) The presence of an extra carboxylate in TMLA could be the reason of the structural flexibility to cause the observed change. An important difference is noticed for kinetic behavior between A and B. While A shows slow reduction and relatively fast oxidation, similar to that found for Zn-TPA MOF, those of B are highly reversible, indicating facile exchange of protons, possibly owing to the presence of hydration water in its structure. Potentiostatic full charging an discharging of the electrodes revealed redox active fractions of 31 and 22% of the total amount of A and B deposited on FTO according to the reactions expressed by Eqs. (A) and (B) above, respectively, being somewhat smaller than the value found for Zn-TPA (35%). However, when the absolute storage capacity per molar amount of Zn is compared, similar values of 1.79 × 10 4 , 1.59 × 10 4 , and 1.69 × 10 4 C mol -1 are found for A, B and Zn-TPA, respectively. Thus, upon improvements of redox active fractions by optimizing the film preparation procedure as well as the thickness, higher storage capacity is expected for Zn-TMLA MOFs than Zn-TPA, because of the contribution of the third carboxylic acid group. Figure 1
    Type of Medium: Online Resource
    ISSN: 2151-2043
    Language: Unknown
    Publisher: The Electrochemical Society
    Publication Date: 2018
    detail.hit.zdb_id: 2438749-6
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 10
    Online Resource
    Online Resource
    The Electrochemical Society ; 2018
    In:  ECS Meeting Abstracts Vol. MA2018-03, No. 4 ( 2018-07-13), p. 233-233
    In: ECS Meeting Abstracts, The Electrochemical Society, Vol. MA2018-03, No. 4 ( 2018-07-13), p. 233-233
    Abstract: Over the past few years, electric power generation from renewable resources such as wind and sun light have already become cheaper than those from fossil fuels and nuclear. However, production of such renewable electricity naturally fluctuates depending on weather. The lack of technically and economically viable systems to store the surplus still hampers the renewables to become primary energy sources. Redox flow batteries (RFBs) are seen as one of the most promising systems of large scale energy storage, because of their separately designable power and capacity, and also its high cycle stability. Vanadium RFBs (VRFBs) are the most established ones, employing redox reactions of vanadium ions in sulfuric acid. Still, high material cost obstructs practical use of VRFBs, especially because proton exchanging Nafion ® membrane takes almost half of the total system cost, followed by 1/4 by vanadium. Alternative redox materials out of cheap elements as well as eliminating Nafion ® are therefore the challenges of top priority. We have recently succeeded in microwave-assisted hydrothermal synthesis of crystalline particles of Zn-based MOFs (metalorganic frameworks) [1]. Double-hydroxide layers of Zn(II) are bridged by terephthalic acid (TPA, benzene-1,4-dicarboxylic acid) to form inorganic/organic hybrid crystals in layered structures, whose spacing and composition vary depending of the pH of precursor solution. The inorganic layer could afford a pathway for electron transport, whereas the interlayer space partially occupied by TPA could allow ion (especially proton) exchange. If a proton-selective redox reaction is achieved, this Zn-TPA MOF could work as a storage electrode for redox batteries in a semi-flow configuration and eliminating membranes. Pastes of Zn-TPA MOFs were prepared by dispersing them into a mixture of 2-butanol and acetylacetone at 35 wt%. The paste was coated on an F-doped SnO 2 (FTO) conductive glass by doctor blading and submerged in warm water to promote necking to fabricate mesoporous electrodes. Electrochemical measurements of the MOF electrodes were performed in a 0.1 M KCl or KI aqueous solution under N 2 , with a Pt wire as counter. Electrode made from Zn-TPA MOF in a Zn 4 (OH) 6 (TPA) composition and with an interlayer distance of 8.18 Å exhibited reversible redox behavior as shown in Fig. 1. Broad cathode peak and relatively sharp anodic peak are observed, both of them being proportional to the square root of the scan rate, but with different slopes. It is most likely that the redox of MOF is coupled with exchange of proton, as expressed by, Zn 4 (OH) 6 (C 8 H 4 O 4 ) + 2H + + 2e - ⇄ Zn 4 (OH) 6 (C 8 H 6 O 4 ) (1) so that diffusion of proton within MOF crystals can be limiting current, and be slow and fast on charging and discharging, respectively. Coulombic reversibility was over 90% at all scan rates and the current was fairly stable over multiple scan cycles. Full charging and discharging under potentiostatic conditions revealed exchange of charge that accounts redox active portion of more than 30% of the total amount of MOF deposited onto the FTO substrate, when assuming the reaction of eq. (1) (1 out of 4 Zn ions undergoes redox). It is therefore clear that the redox reaction is not limited to the Zn ions at the surface of MOFs directly in contact with the electrolyte, but involves those within the crystal bulk, owing to its proton exchanging capability. The same cathodic charging and anodic discharging behavior was observed when the electrolyte was replaced from KCl to KI. On charging, the solution near the counter electrode turned yellow due to formation of I 3 - ions. 3I - + 2e - ⇄ I 3 - (2) Since redox potentials for (1) and (2) are around -1.0 and +0.4 V (vs. Ag/AgCl), respectively, the system could be charged at a voltage of around 1.5 V. Although the cell design as well as the operating condition need to be optimized to achieve a good efficiency, it was possible to maintain some voltage on discharge to prove the system to work as a redox battery. [1] Y. Hirai et al., Microsys. Technol. , 24 , 1, 699–708, 2018. Figure 1
    Type of Medium: Online Resource
    ISSN: 2151-2043
    Language: Unknown
    Publisher: The Electrochemical Society
    Publication Date: 2018
    detail.hit.zdb_id: 2438749-6
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
Close ⊗
This website uses cookies and the analysis tool Matomo. More information can be found here...